Skip to main content

Response mechanisms of Saccharomyces cerevisiae to the stress factors present in lignocellulose hydrolysate and strategies for constructing robust strains

Abstract

Bioconversion of lignocellulosic biomass to biofuels such as bioethanol and high value-added products has attracted great interest in recent decades due to the carbon neutral nature of biomass feedstock. However, there are still many key technical difficulties for the industrial application of biomass bioconversion processes. One of the challenges associated with the microorganism Saccharomyces cerevisiae that is usually used for bioethanol production refers to the inhibition of the yeast by various stress factors. These inhibitive effects seriously restrict the growth and fermentation performance of the strains, resulting in reduced bioethanol production efficiency. Therefore, improving the stress response ability of the strains is of great significance for industrial production of bioethanol. In this article, the response mechanisms of S. cerevisiae to various hydrolysate-derived stress factors including organic acids, furan aldehydes, and phenolic compounds have been reviewed. Organic acids mainly stimulate cells to induce intracellular acidification, furan aldehydes mainly break the intracellular redox balance, and phenolic compounds have a greater effect on membrane homeostasis. These damages lead to inadequate intracellular energy supply and dysregulation of transcription and translation processes, and then activate a series of stress responses. The regulation mechanisms of S. cerevisiae in response to these stress factors are discussed with regard to the cell wall/membrane, energy, amino acids, transcriptional and translational, and redox regulation. The reported key target genes and transcription factors that contribute to the improvement of the strain performance are summarized. Furthermore, the genetic engineering strategies of constructing multilevel defense and eliminating stress effects are discussed in order to provide technical strategies for robust strain construction. It is recommended that robust S. cerevisiae can be constructed with the intervention of metabolic regulation based on the specific stress responses. Rational design with multilevel gene control and intensification of key enzymes can provide good strategies for construction of robust strains.

Graphical Abstract

Background

Carbon neutrality has attracted considerable attention recently due to the increasing CO2 levels in the atmosphere [1]. Biomass is one of the most abundant and renewable resource, which plays an important role to reduce the net CO2 emission [2, 3]. Lignocellulose is a major type of biomass with an estimated yield of hundreds of billions tonnes annually [4], which mainly includes herbaceous straws (e.g., crop straws [5], bagasse [5], leguminosae [6,7,8], solanaceae [9], sunflower [10], etc.), woody biomass (e.g., paulownia tomentosa [11], poplar [12], etc., and various forestry residues), energy crops (e.g., miscanthus and switchgrass [13], etc.) and aquatic biomass. Compared with other biomass such as food crops (corn, soybeans, sugarcane, etc.) [14], and livestock and poultry dung [15]), lignocellulosic biomass has significant advantages like quick production, wide availability, low cost, and without interfering with food security [16]. It has been estimated that plenty of straws with total yield of 3165 million tonnes are produced as agricultural residues every year in the world [5]. Therefore, various technologies have been developed to convert lignocellulosic biomass to biofuels (e.g., bioethanol, butanol, CH4, H2, etc.) and high value-added products (e.g., sugars, alcohols, organic acids, terpenes, etc.) [13, 17].

Lignocellulosic biomass is primarily composed of cellulose (25–55%), hemicellulose (8–50%) and lignin (10–35%) depending on the plant species [18]. Various fuels and chemicals can be produced by thermal, thermochemical and biological conversion of lignocellulosic biomass. However, for bioconversion, the complex structure constructed by the polymeric components necessitates pretreatment of the feedstock by various chemical, physical, and biological methods to deconstruct cell wall structure. Various chemical or combined pretreatments such as dilute acid, alkaline, steam explosion pretreatments, etc., have been developed to facilitate the conversion of carbohydrate polymers to monosaccharide (C5 and C6) [19, 20]. These monosaccharides can be further converted to various biofuels and high value-added platform chemicals such as ethanol, butanol, furfural, sorbitol, etc., through biorefinery technologies (Fig. 1) [13, 17]. However, despite that great efforts have been made to improve cellulose digestibility, cellulosic biofuels and chemicals have yet to be economically feasible at commercial scale [21].

Fig. 1
figure 1

Main stress factors involved in production of lignocellulosic bioethanol by bioconversion

Among the various biomass-derived biofuels, bioethanol is the one with commercialization and annual consumption reaching about 80 million tonnes. The second generation bioethanol with lignocellulose as the feedstock has been considered as one of the most promising way for bioethanol production because of no food-and-fuel debate [22]. The global annual production of cellulosic ethanol would be more than 400 million tonnes if the produced straws are converted to ethanol. However, the production of cellulosic ethanol is still facing various barriers on its way to industrialization. The inhomogeneity and complexity of biomass hydrolysate usually leads to significant decrease in sugar consumption rate and ethanol yield by Saccharomyces cerevisiae. Furthermore, the byproducts formed during pretreatment process, such as organic acids, furan aldehydes, and phenolic compounds, are strong inhibitors to S. cerevisiae growth and metabolism (Fig. 1 and Table 1) [23,24,25]. Thus, one of the obstacles for commercial production of cellulosic ethanol is the lack of robust strains that show strong tolerance to the stress factors [34]. Generally, several issues still need to be solved in order to construct more robust strains. First, most of the engineered strains have good resistance to single stress factor, but the tolerance ability to mixed stress factors are not strong enough [26, 31]. Second, the fermentation efficiency of the engineered strain is still low when lignocellulosic hydrolysate is used as fermentation medium [35, 36]. Third, the endogenous response and regulation mechanisms of S. cerevisiae to various stress factors still needs to be further analyzed and interpreted [37]. With the rapid development of modern biotechnology and synthetic biology, many works have been published for construction of robust strains and revealing the regulatory mechanism of S. cerevisiae in response to stress factors [38,39,40,41,42]. The response of S. cerevisiae has significant difference to various stress factors [24]. Although there are some similarities in different responsive metabolic pathways at the cellular level, the functional genes involved in can be much different [27, 43]. Considering the complexity of the metabolic regulation of S. cerevisiae and the certain gap for large-scale industrial applications, in this work, we have reviewed the stress response mechanisms of S. cerevisiae to various stress factors and the potential targeted genes. The strategies to improve strains’ tolerance ability also have been discussed, which may provide technical inspiration for construction of robust strains for industrial production of bioethanol with high efficiency.

Table 1 Recently reported fermentation performances of S. cerevisiae with lignocellulosic hydrolysate as the carbon

Response of S. cerevisiae to stress factors

Pretreatment is a pre-requisite step to facilitate the release of fermentable sugars either by chemical or enzymatic hydrolysis. Dilute acid pretreatment has been considered as one of the most promising approaches with potential commercial applications because the process employs cheap mineral acids such as sulfuric acid and can hydrolyze most of hemicelluloses to fermentable sugars [44]. However, stress factors including organic acids, furan aldehydes, and phenols are formed inevitably [45] (Fig. 1). The types and contents of these stress factors are closely related to the pretreatment conditions [46, 47]. Severe pretreatment conditions can promote the release of fermentable sugars, but also lead to formation of more stress factors. The yeast usually can give positive stress response to maintain its growth and reproduction in the presence of these stress factors [48]. However, there is a critical level of the stress factors concentration for yeast to initiate stress response. The growth and fermentation performance of the yeast can be greatly affected when the concentrations of the stress factors exceed the critical levels [23, 24].

Organic acids

Organic acids stress factors mainly refer to formic acid and acetic acid. Acetic acid is mainly formed by deacetylation of the acetyl group of hemicelluloses, while formic acid can be formed by degradation of sugars [49]. The concentrations of organic acid stress factors in the hydrolysate are relatively higher than furans and phenols. The concentration of acetic acid is 1–15 g/L depending on the biomass feedstock and pretreatment methods used [50, 51]. The concentration of formic acid is usually lower than that of acetic acid, but the inhibitory effect of formic acid is stronger because of its lower pKa value (3.75 for formic acid in contrast to 4.75 of acetic acid). Intracellular acidification is the main reason for the inhibitive effects of organic acids on the cell growth [23, 24]. To maintain intracellular pH homeostasis, H+ is pumped out of the cell with the help of the ATPase in the membrane. A large amount of ATP is thus consumed with the exhaustion of H+, which leads to insufficient intracellular energy supply and affects cell growth and metabolism. At the same time, the extracellular organic acids could be transferred into cell. The intracellular organic acids continuously dissociate to release H+ for maintaining intracellular and extracellular ion balance. Since anions cannot be pumped out, the accumulated anions in the intracellular seriously affect the performance of cells [52].

Furan aldehydes

Two types of furan aldehydes, namely furfural and 5-hydroxymethylfurfural (HMF), are usually detected in the hydrolysate obtained by thermochemical pretreatment of lignocellulosic biomass due to the dehydration reactions of pentoses and hexoses [53, 54]. Furfural concentrations usually ranges from 0.5 to 3 g/L, while the hydrolysate from corn stover can contain furfural as high as 11 g/L [55]. HMF usually found in spruce hydrolysate with the concentrations varied from 2.0 g/L to 5.9 g/L [56]. Intracellularly, furfural is reduced to furfuryl alcohol by alcohol dehydrogenases (ADH) or reductases (AKR/ARI) with NAD(P)H as a cofactor. Furfuryl alcohol could be rapidly converted by alcohol dehydrogenase [57, 58]. Furfural can also be oxidized to its acid form (furoic acid) by aldehyde oxidase, and then transformed into 2-oxoglutaric acid by a series enzyme, and further metabolized via tricarboxylic acid cycle (TCA) [24, 57]. The mechanism of intracellular detoxification of HMF is similar to that of furfural. However, due to the poorer ability of HMF to penetrate cell membrane than that of furfural, it takes longer time for the cells to detoxicate HMF [24, 59]. Furfural and HMF can be oxidized and reduced to lower-toxic substances in cells. However, this detoxification process breaks the intracellular redox balance. As a result, acetaldehyde accumulates during reduction of furfural/HMF, which is most likely to contribute to the inhibition of cell growth [23, 24].

Furan aldehydes have strong toxicity to S. cerevisiae [60]. The yeast showed a prolonged lag phase of 8 and 4 h for furfural- and HMF-treated cultures, respectively. The inhibition completely suppressed the cell growth at 120 mM of furfural or HMF independently. However, when both inhibitors were co-present in the media, cell growth was only recovered at 10 mM of each inhibitor, indicating that these inhibitors acted in negative synergic fashion even at low concentrations [61]. The activity of enzymes (hexokinase, triosephosphate dehydrogenase, and alcohol dehydrogenase) involved in glycolysis could be significantly inhibited under 10 mM furfural stress [62]. The genes related to oxidative stress response are upregulated in S. cerevisiae under furfural stress. Furfural can induce the accumulation of reactive oxygen species in cell, resulting in the irreversible damage to mitochondria and vacuole membrane, actin, cytoskeleton, and nuclear chromatin [63]. Furthermore, furfural and HMF can inhibit the protein and RNA synthesis, and reduce enzymatic and biological activities [61]. Therefore, although the concentrations of furan derivatives are usually lower than that of organic acids, they show more serious inhibitory effects on the growth and metabolism of yeasts.

Phenolic compounds

Phenolic compounds present in the hydrolysates are typically formed by degradation of lignin, and usually exist in four forms, namely phenolic acids (e.g., ferulic acid), phenolic aldehydes (e.g., vanillin), phenolic ketones (e.g., 4-hydroxyacetophenone), and phenolic alcohols (e.g., homovanillyl alcohol) [64]. However, detailed characterization of the formed phenolic compounds is challenging because of their diversity and complexity in structure. The inhibition mechanism of these compounds might be attributed to cytoplasmic membrane invaginations, decreased membrane potential and permeability, repressed translation process, inhibited ribosomal function, and increased intracellular ROS concentration [65,66,67,68]. Phenolic compounds have serious stress effects on the growth and reproduction of S. cerevisiae, in which phenolic aldehydes shows the strongest toxicity. Coniferyl aldehyde even at a low concentration (1 mM) has been found to completely inhibit cell growth [69]. Phenolic compounds not only inhibit the cell growth, but also lead to deactivation of cellulases, resulting in the decrease in the yield of fermentable sugars by subsequent enzymatic hydrolysis [70, 71]. Therefore, improving the tolerance of strains to phenolic compounds is of great significance for bioethanol production from lignocellulose biomass.

Synergetic effects of the stress factors

Various stress factors coexist in the hydrolysate, and thus the synergetic effects of these inhibitors exert more serious stress on S. cerevisiae [31]. Compared with single stress factor, mixed stress factors may induce more complex response with consumption of more intracellular energy resulting in a significant decrease in fermentation efficiency. However, the interactions among various stress factors, including positive and negative synergistic effects, still need further interpretation [72]. Chen et al. [43] reported that the glucose could be depleted in 60 h and 40 h under 2 g/L acetic acid and 1.5 g/L furfural stress, respectively, while the glucose was not exhausted even fermentation for 96 h under mixed acetic acid and furfural (2 + 1.5 g/L) stress. Under the stress effect of these factors, transmembrane transport processes play pivotal roles in response to acetic acid, and carbohydrate metabolic process is crucial for furfural stress. While the biological processes such as transmembrane transport, cellular amino acid metabolic process, and response to inhibitors are involved in S. cerevisiae resistance to the mixed fermentation inhibitors. Li et al. [41] revealed that more differentially expressed genes (DEGs) were involved in S. cerevisiae response to mixed acetic and formic acid stress compared with that to only acetic acid or formic acid stress. There were 294 unique DEGs among 657 total DEGs under the mixed acetic and formic acid stress, indicating that 45% of DEGs were unique in response to the mixed acid stress and 55% of DEGs were shared to respond single acid stress. It indicates that the strain adopts distinct regulatory mechanisms to reprogram cell metabolism in response to various stress factors.

Many studies have focused on revealing the stress response of S. cerevisiae to mixed stress factors with synthetic medium. However, when lignocellulose hydrolysate is used as fermentation medium, the stress response mechanisms of strains becomes more complicated which still need to be further interpreted [39, 73]. Revealing the stress response of S. cerevisiae induced by various stress factors is of great significance for developing robust yeasts for industrial purpose, since new strategies can be made by intensifying the key genes to detoxify the inhibitors.

Response mechanisms of S. cerevisiae to stress factors

The response of S. cerevisiae to stress factors often presents global control at the cell level. With the continuous development of molecular biology and genetic engineering technology, many studies have focused on transcriptome and proteomic analysis to accurately and clearly reveal the endogenous regulation mechanisms of S. cerevisiae [42]. Based on these revealed results, the potential response mechanisms of S. cerevisiae have been summarized from cell wall/membrane regulation, energy regulation, amino acid regulation, transcriptional and translational regulation, and redox regulation, as shown in Fig. 2.

Fig. 2
figure 2

Endogenous regulation mechanisms of S. cerevisiae in response to stress factors

Cell wall/membrane regulation

Cell wall/membrane is the first barrier of S. cerevisiae to resistant to stress factors [74,75,76]. The integrity and selective permeability of the cell wall/membrane are very important for strains to cope with external stress. The cell wall of S. cerevisiae contains four classes of macromolecules, namely cell wall proteins (CWPs), β-1,6-glucan, β-1,3-glucan, and chitin [77, 78]. These compositions and their degrees of cross-link can change in response to varied stress conditions [79, 80]. Cell wall is able to transmit signals to the cell membrane under stress, which remodels the components and structures of the membrane to adapt the corresponding stress [81]. The genes encoding these substances are increased to maintain the cell wall homeostasis in the presence of stress factors. The cell wall- and membrane-related genes, including AUS1, COX14, GSC2, INP52, MCH2, MDG1, MDY2, NAB6, PEX3, PMT1, PSA1, SED1, SPT14, TIR4, USV1, and YTP1 have been found to upregulate in response to furfural and HMF stress [79]. Other genes that may not necessarily locate in cell wall but encode cell wall-related proteins are also involved in cell wall homeostasis in response to osmotic pressure, high temperature, furan aldehydes, or other stress factors, such as CWP1, PIR3, PKC1, RHO1, ROM1, RLM1, SLT2, and YGP1, related to cell wall integrity (CWI) signaling pathway; PIR3, SPT14, SED1, SPI1, and TIR4 encode glycosylphosphatidylinositol-linked cell wall proteins; DIT1, GIP1, and GSC2 for spore wall formation; PSA1 and USV1 for cell wall biosynthesis; YPS3 for cell wall growth and maintenance; and PMT1 for cell wall rigidity [82,83,84]. The genes involved in membrane function are also upregulated in the presence of stress, such as AUS1, PDR15, and YOR1 involved in plasma membrane ATP-binding cassette (ABC) transporter, HSP12 involved in maintaining membrane organization [76, 79]. These alteration results in cells that are more resistant to further cell wall degradation. Therefore, systemically regulating the expression of these genes to strengthen the composition and function of the cell barrier can effectively improve the stress resistance of strain.

Energy regulation

Central carbon metabolism (CCM) is the main source of energy required for cell metabolism and provides precursors for other metabolites, which mainly includes glycolysis/gluconeogenesis, pentose phosphate pathway, and tricarboxylic acid cycle [85]. CCM has been found to play key roles in modulating yeast survival in response to stress factors. The genes involved in glycolysis are generally upregulated in response to 40 mM acetic acid [41], while the genes are downregulated when the acetic acid concentration reaches 300 mM [86]. Guo et al. [87] indicated that organic acids at moderate concentrations could stimulate the glycolytic flux, while higher acid level slowed down the glycolytic flux for both aerobic and anaerobic growth of S. cerevisiae. Chen et al. [43] found that carbohydrate metabolic process was crucial for strain response to furfural. The expression of most glycolytic enzymes are increased in industrial S. cerevisiae, whereas those in the TCA cycle, glycogen and glycerol biosynthesis, and pentose phosphate pathway are largely downregulated in response to thermal stress [88].

Some other carbohydrates participating in CCM through oxidation–reduction reactions are also differentially expressed in response to stress factors. These substances play important roles in strain defense system. Trehalose metabolism (TPS2, TSL1, and ATH1) has been found to protect cell biomacromolecules from stress effects caused by acetic acid, high temperature, and high osmotic pressure [86, 89]. Glucan synthesis (FKS1, FKS2, and ROM2) has been found to involve in cell wall remodeling under acetic acid stress [90, 91]. The other carbohydrates, such as inositol phosphate and galactose, also play indelible roles in strains’ response to stress effects [41].

Amino acid regulation

Amino acids are the key hallmarks and mediators for S. cerevisiae in response to stress factors, which essentially serve as a nitrogen source and the building blocks of proteins [92,93,94]. Increasing the expression of amino acid metabolism-related target genes or transcription factors (TFs) have been reported to contribute to the improving of the strain tolerance [27]. The genes involved in arginine, histidine, and tryptophan, were upregulated in response to acetic acid [86]. Tryptophan has a prominent contribution to maintaining cell membrane stability in S. cerevisiae [95], which is beneficial to enhance the tolerance of strain to stress factors. The genes related to the biosynthesis of cysteine and methionine (CYS3 and MET4), histidine (HIS4), glycine (GLY1), and glutamate (GDH1) were identified as determinants of resistance to acetic acid [96]. Supplementation of cysteine, glycine, and glutamate (20 mg/L for each amino acid) could slightly increase S. cerevisiae resistance to acetic acid [96]. Except for these amino acids, alanine, aspartate, serine, threonine, proline, phenylalanine, tyrosine, valine, leucine, isoleucine, and beta-alanine metabolism are also involved in stress resistance [27, 97].

Transcriptional and translational regulation

Genetic information flows from DNA to RNA, and then translates into proteins. This process is directly involved in the regulation of the growth and reproduction of strain, as well as in the anabolism and catabolism of key substances. S. cerevisiae must dynamically alter the levels of transcription and translation to respond the diverse stress [98]. The ADEnine requiring (ADE) genes, including ADE1, ADE13, and ADE17, etc., participate in the de novo purine biosynthesis pathway yielding inosine monophosphate (IMP) and adenosine 5′-monophosphate (AMP) [99]. Overexpressing of these ADE genes in S. cerevisiae have been found to enhance cell growth and ethanol productivity under mixed acetic acid, formic acid, furfural, and 5-HMF stress [100]. The genes involved in mRNA synthesis that directly affect transcriptional control and RNA process are downregulated in response to furfural, and the genes involved in mitochondrial ribosomal proteins are downregulated in response to acetic acid [86]. High concentrations of vanillin result in the repression of translational and the formation of cytosolic mRNP granules, leading to a reduction in overall protein synthesis levels and the limited translation of mRNAs [68, 101,102,103]. Regulating the expression of these genes to relieve stress-induced transcriptional and translational repression can be beneficial to enhance the tolerance of strain.

Redox regulation

Redox regulation is a universal response of S. cerevisiae to resistant various stress factors. The oxidant defense systems mainly include NAD(P)+/NAD(P)H homeostasis, superoxide dismutase (SOD), reduced thioredoxin, glutathione reductase, and vitamin C [97, 104, 105]. NAD(P)/NAD(P)H homeostasis, correlated to NAD+ synthesis, and redox transformation from NAD(P)+ to NAD(P)H, is essential for preventing intracellular acidification induced by weak acid, and acidified phenolic or furan compound [97]. Various stress factors can induce the reactive oxygen species (ROS) accumulation in cell. SOD protects cells by scavenging ·O2 (ROS). It has been reported that overexpression of SET5 and PPR1 or deletion of ADY2 and JJJ1 are beneficial to decrease the ROS accumulation and endow yeast increased tolerant ability to acetic acid [106, 107]. The nonenzymatic defense systems including reduced thioredoxin, glutathione reductase, and vitamin C, can act as reducing agents for scavenging free radicals to maintain intracellular redox homeostasis [108, 109].

Target genes for improving S. cerevisiae tolerance to stress factors

Two states, activation and inhibition, are exhibited by yeast strains in response to stress factors, and ultimately are reflected by the upregulation or downregulation of genes. Based on the development of transcriptomics, proteomics, metabolomics, and other omics technologies, numerous potential targeted genes that have the positive contribution or regulatory function under stress conditions have been revealed through experiments (Tables 2 and 3). Due to the complexity of the endogenous regulatory mechanisms of strain and the lack of systematic understanding for the function of the targeted genes, the improvement of strain performance is very limited [141]. The genes and TFs that have been revealed to have contribution to improving the strain tolerance are summarized as shown in Tables 2 and 3.

Table 2 Reported target genes in improving S. cerevisiae resistance to stress factors
Table 3 Reported target TFs in improving S. cerevisiae resistance to stress factors

Potential function genes involved in strain defense

The enzymes encoded by target genes is involved in the key signaling pathways and substance metabolism pathways, which contribute to the resistance ability of strains in response to specific stress. The experimentally proven functional genes are mainly involved in cell wall barrier (CWB), cell membrane barrier (CMB), basic metabolism maintenance (BMM), and intracellular redox homeostasis (IRH) (Table 2). These functional modules are closely correlated with the response mechanisms of S. cerevisiae (Fig. 2). When the cell wall receives stress signals, the protein Mid2p or Wsc1/2/3p (cell surface sensors) will be activated and then stimulate Rom2p (guanine nucleotide exchange factor) to activate the expression of Rho1p (small GTPase). The protein Rho1p that is involved in the establishment of cell polarity, regulates the expression of Pkc1p (protein kinase C). Pkc1p is essential for cell wall remodeling during growth, and in turn stimulates the cell wall integrity pathway. The PKC1-mediated signaling pathway further regulates the expression of Slt2p that is involved in the maintenance of cell wall integrity [74]. The activation of Slt2p results in changed transcription of more than twenty genes related to cell wall (Fig. 3). These genes cooperate to maintain the cell wall homeostasis [83]. This pattern of gene interaction is also presented in carbohydrate metabolism, amino acid metabolism, and intracellular redox regulation (Fig. 3 and Table 2). The key substances, glycerate-3-phosphate, pyruvate, oxaloacetic acid, ketoglutaric acid, and erythrose-4-phosphate, in central carbon metabolism are also involved in amino acid metabolism through enzymatic catalysis. Regulating the expression of the key genes in pentose phosphate pathway (TAL1, TKL1, GND1, and ZWF1) not only regulates energy metabolism and basic substance metabolism, but also helps maintain intracellular NADP+/NADPH content that participates in the intracellular redox regulation [117, 125, 128].

Fig. 3
figure 3

Schematic diagram of the multilevel defense of S. cerevisiae to different stress factors, including cell wall signals cascade, plasma membrane barrier, central carbon and amino acid metabolism, and intracellular redox homeostasis

This complex regulatory network is related to the fact that the improvement of the strain performance depending on single gene positive regulation is not enough to significantly improve the fermentation efficiency in the presence of stress factors. The up-/down-regulation of the key genes results in the accumulation or deprivation of the downstream/upstream products, which in turn leads to the disequilibrium of metabolic networks (such as feedback regulation). Therefore, assisted domestication or mutagenesis may be a method to balance cell metabolic homeostasis on the basis of limited improved strain performance by directional modification [142].

Potential transcription factors involved in strain defense

TFs can regulate a series of genes. The disturbance caused by the up-/down-regulation of a single TF may be more significant than that caused by a single gene. However, the performance of the strain obtained by directional modification of single TF still cannot reach the ideal level when lignocellulosic hydrolysate is used as fermentation medium to produce ethanol [31, 143]. This is not only related to the complex fermentation environment of hydrolysate, but also related to the unclear regulatory mechanism of TFs. Therefore, the regulation mechanisms of the typical TFs are briefly summarized (Table 3).

Haa1p regulation

Haa1p is involved in the endogenous regulation of S. cerevisiae and the reduction of intracellular acetate concentration under acetic acid stress [144]. Except to acetic acid, Haa1p is also related with yeast resistance to lactic acid and propionic acid [131, 145]. Haa1p rapidly relocates from the cytoplasm to the nucleus in the presence of acetic acid or lactic acid [146, 147]. The DNA binding of Haa1p can be induced by acetic acid. Acetate binds to the N-terminal 150-residue region (N-terminal Zn-binding domain), and the transcriptional activation domain is located between amino acid residues 230 and 483 [148]. Haa1p involved in adaptation to weak acid stress by inducing the transcription of genes TPO2 and TPO3, which localize on the plasma membrane and encodes polyamine transporter of the major facilitator superfamily [131, 149]. Under the acetic acid stress, Haa1p also regulates the genes that encode protein kinases, multidrug resistance transporters, membrane transporter, membrane stress proteins [144, 150]. The genes involved in lipid metabolism, and nucleic acid processing are also directly or indirectly regulated by Haa1p in response to acetic acid [144]. The strain overexpressed HAA1 exhibits improved sugar consumption and ethanol production from glucose or xylose in the presence of acetic acid [11, 151]. These conclusions can prove that Haa1p has a prominent contribution to improving the tolerance of S. cerevisiae to weak acid in hydrolysates, especially acetic acid.

Msn2/4p regulation

Msn2p and Msn4p are two homologous stress-responsive TFs involved in S. cerevisiae transcriptional response to environmental stress response, such as acetic acid, furfural, oxidative stress, osmotic shock, glucose starvation, high ethanol concentrations, and high temperature [92, 152]. Once S. cerevisiae is challenged by these stress factors, Msn2/4 are rapidly dephosphorylated and translocated into the nucleus [153, 154]. Msn2/4p binds DNA at stress response elements of responsive genes and activates hundreds of stress-related genes as a consequence to various stress conditions [155]. The reported functional domains of MSN2 include the C-terminal zinc finger DNA-binding domain (DBD), the nuclear localization signal (NLS) region, the nuclear export signal (NES) region, and the imperative transcriptional activating domain (TAD) at the N-terminus [154, 156,157,158]. The genes encoding antioxidant enzymes (CTT1, SOD1, SOD2, PRX1, and TSA2) were regulated by MSN2/4, which is beneficial for removing reactive oxygen species (ROS) to eliminate stress from various stress factors [104, 135]. The stress-related genes regulated by MSN2/4 also involved in protein quality control (HSP12, HSP26, HSP42, HSP82, HSP104, SSA1, and SSA4), mitochondrial respiratory (COX5b, COX17, and COX20), glycogen synthetic (GSY1, GSY2, and GLC3), and pentose phosphate pathway (SOL4, GND2, and TKL2) [135, 159, 160]. Previous studies have indicated that the transcription abundance of MSN2/4 could be significantly increased in response to acetic acid, furfural, or high temperature, etc. [79, 134]. These facts demonstrate that overexpression of MSN2/4 is a promising approach for constructing robust strains with improved tolerance and fermentation performance.

Yap1p regulation

The basic leucine-zipper transcription factor Yap1p transits from the cytoplasm to the nucleus when triggered by oxidative stress, and is degraded in the nucleus after the oxidative stress has removed [143]. Cysteine-rich domain is the active site of Yap1p [137]. The specific activity of enzymes involved in oxidative detoxification, such as glucose-6-phosphate dehydrogenase, superoxide dismutase, and glutathione reductase could be decreased with the knockout of Yap1p, which increases the sensitivity of yeast cell to hydrogen peroxide and chemicals that generate superoxide anion radicals [161, 162]. In addition to oxidative stress, furfural and HMF acting as thiol-reactive electrophiles could directly activate Yap1p. Overexpression of YAP1 enhances the tolerance of S. cerevisiae to furfural and HMF by activating catalase expression (CTA1 and CTT1) [120, 140, 163].

Based on omics analysis and various molecular techniques, more tolerant-related TFs and their regulatory mechanisms are being revealed, providing biological basis for robust strain construction with rational design. Haa1p, Msn2/4p, and Yap1p are typical TFs in improving the tolerance of S. cerevisiae. Other TFs, such as Skn7p (involved in oxidative stress response) [137], Sfp1p and Ace2p (involved in acetic acid and furfural stress response) [43] are also involved in the stress response of S. cerevisiae. In addition to these TFs verified by genetic engineering, transcriptome and other omics analysis also have revealed some potential TFs that contribute to stress response, such as Hcm1p, Fkh1/2p, Pdr1/3p, Met4p, etc. [52, 120, 164]. Tolerance-related TFs are synergistically involved in strain stress response. Current experimental results have confirmed that the performance of the engineered strains can be effectively improved by regulating key genes or TFs in the laboratory stage, but the stress resistance of the strains still needs to be improved in industrial applications [26]. It once again suggests that the improvement of strain robustness should be based on synergetic regulation of multi-genes/pathways. Therefore, understanding the response mechanism of S. cerevisiae to various stress factors and the key genes or TFs, is of great significance for constructing robust strains.

Strategies for construction of robust strains

The interaction of multi-genes/pathways for stress-induced response and the endogenous regulation mechanisms of S. cerevisiae on stress factors are schematically shown in Figs. 2 and 3. The complex endogenous regulatory mechanisms of the yeast, as well as the fermentation performance and tolerability of the engineered strains suggest that the positive regulation of a single gene may be unable to cope with the cell-level disturbances. The mixed stress factors in the fermentation system puts forward higher requirement for the robustness of the yeast. Therefore, synergetic regulation of multigene to establish multi-tolerant system may be one of the most efficient methods to improve the fermentation performance of the yeasts in response to stress factors (Fig. 4). As a result, researchers have proposed to build a robust system based on genetic engineering (such as, synthetic chromosome rearrangement and modification by LoxP-mediated evolution (SCRaMbLE) [165,166,167], global transcription machinery engineering (gTME) [168], RNA interference (RNAi)-assisted genome evolution (RAGE) [169], automated multiplex genome-scale engineering (CRISPR–Cas) [170]), which can induce the rewriting of the endogenous regulation to establish tolerance response. At the same time, domestication and mutagenesis (such as, atmospheric and room temperature plasma (ARTP) and UV mutagenesis) is assisted to balance the metabolic load of strain [171,172,173]. Furthermore, targeted elimination of stress factors such as weak acids, furans, and phenols, has been proposed to cope with the stress from hydrolysate components, considering that these substances can be converted to less toxic substances by specific enzymes [50, 57]. This strategy also relies on the homologous or heterologous expression of the target genes. Assisting adaptive evolution will be more conducive to the target strain construction with the disturbance of the endogenous regulation caused by the target gene expression.

Fig. 4
figure 4

Possible strategies to construct robust S. cerevisiae. The response performance of S. cerevisiae to stress factors can be improved with the construction of the tolerant and eliminated pathway in cell. The expression of the single/multiple genes circuit induces the rewriting of the endogenous regulation of S. cerevisiae. Assisting in adaptive evolution to balance the metabolic load of strain can further promote the improvement of S. cerevisiae robustness

Single-site modification

At present, many works have been done to endow S. cerevisiae with tolerance to certain stress factors through directional modification of the up-/down-regulation of the key genes (Tables 2 and 3). However, this strategy just can increase the strain robustness with limited improvement. A common problem has been encountered is that the engineered yeasts still show relatively low fermentation efficiency when lignocellulose hydrolysate is used as the carbon source (Table 1). This is because the co-presence of various stress factors in hydrolysate causing more serious stress on the growth of the yeast.

Multi-site modification

Multi-site modification based on rational and irrational genetic editing techniques may be one of the effective approaches to obtain robust strains [174]. Rational construction mainly relies on the directional design of functional genes to construct multiple defense system. Irrational construction can be realized through domestication, ARTP, and other technical means to achieve non-directional change of multiple loci, so as to establish a mutant yeast library, and obtain excellent strains through screening and evaluation. Xu et al. [142] developed a multilevel defense system (MDS) by random assembly of tolerance genetic circuits, adaptive evolution (e.g., ARTP), and multi-step screening to obtain industrial yeasts with higher robustness and productivity. The strain integrated with MDS can tolerant multiple stress such as high surge, high temperature, and ethanol. The fermentation performance was significantly improved compared with single defense system. Si et al. [169] reported the RAGE technology as a generally applicable method for genome-scale engineering in S. cerevisiae. RNAi-assisted genome evolution could improve the acetic acid tolerance of strain. Except for the application of RAGE, Si et al. [170] developed a platform for automated multiplex genome-scale engineering in S. cerevisiae, which iteratively integrated the functional gene fragment of cDNA library into the genome with the aid of CRISPR–Cas. This system allowed functional mapping and multiplex optimization on a genome scale for diverse phenotypes, such as acetic acid tolerance.

Multilevel directional modification can be obtained by integration of multiple target genes/fragments with specific functions, in which designing and revealing the target genes/fragments with superior performance is the basis for successful construction of robust strain. Omics data analysis have indicated that cell wall/membrane integrity, energy metabolism, amino acid metabolism (protein quality control), and redox homeostasis play an important role in strain resistance (Figs. 2 and 3). Therefore, it is expected to construct robust S. cerevisiae strain by regulating the expression of these genes. Meanwhile, assisting domestication, mutagenesis, and other techniques to adjust the degree of adaptation between the genetic circuits and chassis, can further improve the robustness of the strain.

Metabolism and elimination of stress factors

Improving the ability of strain to eliminate stress factors (e.g., organic acids, furan aldehydes, and phenols) is another tolerant feature of the robust S. cerevisiae. The up-take of these stress factors enables in situ detoxification of lignocellulosic hydrolyzates to lessen their inhibitory effects. Weak acids enter yeast cells by free diffusion and active transportation [175]. By combining a nicotinamide adenine dinucleotide (NADH)-consuming acetate consumption pathway and an NADH-producing xylose utilization pathway, Wei et al. [50] successfully constructed a pathway to convert toxic acetic acid to ethanol in engineered S. cerevisiae under anaerobic conditions. Furfural and HMF can be used as the sole carbon source for cell growth by Amorphotheca resinae ZN1, and converted to low-toxic compounds for S. cerevisiae [57, 122]. Overexpression of dehydrogenases (ADH6/7) and pentose phosphate pathway (ZWF1) genes can increase the reduction capacity of S. cerevisiae to furfural and 5-HMF [119, 122, 123]. For the elimination of phenolic compounds, the reported studies mainly rely on laccase to degrade it into low-toxic compounds outside the cells. The growth rate and ethanol productivity of S. cerevisiae were increased with the laccase treatment for the phenolic compounds [176]. Lei et al. [115] integrated laccase gene (lacA, from Trametes sp. AH28-2) that fused with a-factor signal sequence into S. cerevisiae CEN.PK, and further overexpressed chaperone gene (KAR2) to promote the translocation of laccase. Their results showed that vanillin-specific conversion rate was increased, and the strain tolerance to vanillin was increased.

The above-mentioned construction strategies can be achieved by the upregulation, downregulation, or knockout of the target genes and TFs. Therefore, systematic understanding the endogenous regulatory mechanisms of strain can lay a foundation for robust strain construction for efficiently conversion of lignocellulosic hydrolysate to biofuels and chemicals.

Conclusions and prospects

Bioconversion of lignocellulosic biomass to bioethanol is one of the important ways to achieve carbon neutrality. However, relatively low tolerance of S. cerevisiae to the stress factors in the hydrolysate is one of the key obstacles for direct utilization of lignocellulosic hydrolysate. S. cerevisiae displays somewhat different response to different stress factors, such as organic acids that mainly cause intracellular acidification, furan aldehydes that induce intracellular redox imbalance, and phenolic compounds that destroy the cell membrane integrity. These negative effects can induce the rewriting of the endogenous regulatory of S. cerevisiae by differential expression of a series of genes mainly regarding the cell wall/membrane, energy, amino acids, transcriptional and translational, and redox regulation. Based on the above response mechanism, the robustness of S. cerevisiae might be improved by designing target genetic circuit according to corresponding regulatory points. Due to the complex regulatory network of S. cerevisiae, simple regulation of several genes or TFs usually cannot meet the requirements to obtain enough high robustness of strains for industrial application. Many strategies including single-site modification, multi-site modification, and metabolic intensification to improve strain tolerance and eliminate stress factors have been developed for construction of robust strains owing to the rapid development of modern biotechnology and synthetic biology such as SCRaMbLE, RAGE, gTME, CRISPR–Cas, adaptive evolution, multi-step screening, etc. With the aim of further increasing the tolerance of S. cerevisiae, various strategies can be employed by effective integration of the target genetic circuit. Therefore, future works can be done at least from following aspects.

(a) Diverse gene circuits for improving the strain robustness could be designed, including cell barriers, energy supply pathways, antioxidant defense, identification, and elimination systems. Combined gene circuits may be obtained by integrating functional modules with better tolerance.

(b) Construction of engineered cells with different tolerance preferences using the same parent strain, respectively, to reduce the metabolic burden of multigene integration. These cells then can be co-cultured to eliminate various stress factors in hydrolysate. Therefore, the concept of “functional flora” can be applied to cope with the complex stress of multi-stress factors by using several independent cells (the same parent strain) with different functions.

(c) Combining directional and non-directional modification techniques to balance the metabolic burden of the robust strains to tolerant multi-stress factors. The strain performance can be improved by target genetic modification. By assistance of domestication, ARTP, and other mutagenesis technologies, it is expected to adjust the metabolic load of strains and further improve the performance.

Availability of data and materials

Not applicable.

Abbreviations

ABC:

ATP-binding cassette

ADH:

Furfuryl alcohol dehydrogenases

AMP:

Adenosine 5′-monophosphate

ARTP:

Atmospheric and room temperature plasma

BMM:

Basic metabolism maintenance

CCM:

Central carbon metabolism

CMB:

Cell membrane barrier

CWB:

Cell wall barrier

CWI:

Cell wall integrity

CWP:

Cell wall protein

DBD:

DNA-binding domain

DEG:

Differentially expressed genes

gTME:

Global transcription machinery engineering

HMF:

5-Hydroxymethylfurfural

IMP:

Inosine monophosphate

IRH:

Intracellular redox homeostasis

MDS:

Multilevel defense system

NES:

Nuclear export signal

NLS:

Nuclear localization signal

RAGE:

RNA interference (RNAi)-assisted genome evolution

ROS:

Reactive oxygen species

SOD:

Superoxide dismutase

TAD:

Transcriptional activating domain

TCA:

Tricarboxylic acid cycle

References

  1. Liu Z, Wang K, Chen Y, Tan T, Nielsen J. Third-generation bio-refineries as the means to produce fuels and chemicals from CO2. Nat Catal. 2020;3(3):274–88.

    Article  CAS  Google Scholar 

  2. Gustavsson L, Börjesson P, Johansson B, Svenningsson P. Reducing CO2 emissions by substituting biomass for fossil fuels. Energy. 1995;20(11):1097–113.

    Article  CAS  Google Scholar 

  3. Wu Y, Zhao F, Liu S, Wang L, Qiu L, Alexandrov G, Jothiprakash V. Bioenergy production and environmental impacts. Geosci Lett. 2018;5(14):1–9.

    CAS  Google Scholar 

  4. Dahmen N, Lewandowski I, Zibek S, Weidtmann A. Integrated lignocellulosic value chains in a growing bioeconomy: Status quo and perspectives. GCB Bioenergy. 2019;11(1):107–17.

    Article  Google Scholar 

  5. Tan J, Li Y, Tan X, Wu H, Li H, Yang S. Advances in pretreatment of straw biomass for sugar production. Front Chem. 2021;9:696030.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Turan V. Arbuscular mycorrhizal fungi and pistachio husk biochar combination reduces Ni distribution in mungbean plant and improves plant antioxidants and soil enzymes. Physiol Plant. 2021;173(1):418–29.

    CAS  PubMed  Google Scholar 

  7. Turan V, Schröder P, Bilen S, Insam H, Juárez M. Co-inoculation effect of rhizobium and Achillea millefolium L. oil extracts on growth of common bean (Phaseolus vulgaris L.) and soil microbial-chemical properties. Sci Rep. 2019;9(1):15178.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  8. Akcura M, Turan V, Kokten K, Kaplan M. Fatty acid and some micro element compositions of cluster bean (Cyamopsis tetragonoloba) genotype seeds growing under Mediterranean climate. Ind Crops Prod. 2019;128:140–6.

    Article  CAS  Google Scholar 

  9. Turan V, Khan SA, Iqbal M, Ramzani PMA, Fatima M. Promoting the productivity and quality of brinjal aligned with heavy metals immobilization in a wastewater irrigated heavy metal polluted soil with biochar and chitosan. Ecotoxicol Environ Saf. 2018;161:409–19.

    Article  CAS  PubMed  Google Scholar 

  10. Turan V, Ramzani PMA, Ali Q, Abbas F, Iqbal M, Irum A, Khan WUD. Alleviation of nickel toxicity and an improvement in zinc bioavailability in sunflower seed with chitosan and biochar application in pH adjusted nickel contaminated soil. Arch Agron Soil Sci. 2018;64(8):1053–67.

    Article  CAS  Google Scholar 

  11. Cunha JT, Costa CE, Ferraz L, Romaní A, Johansson B, Sá-Correia I, Domingues L. HAA1 and PRS3 overexpression boosts yeast tolerance towards acetic acid improving xylose or glucose consumption: unravelling the underlying mechanisms. Appl microbiol biot. 2018;102(10):4589–600.

    Article  CAS  Google Scholar 

  12. Wang ZJ, Zhu JY, Zalesnyjr RS, Chen KF. Ethanol production from poplar wood through enzymatic saccharification and fermentation by dilute acid and sporl pretreatments. Fuel. 2012;95:606–14.

    Article  CAS  Google Scholar 

  13. Rosales-Calderon O, Arantes V. A review on commercial-scale high-value products that can be produced alongside cellulosic ethanol. Biotechnol Biofuels. 2019;12:240.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  14. Barr MR, Volpe R, Kandiyoti R. Liquid biofuels from food crops in transportation-a balance sheet of outcomes. Chem Eng Sci X. 2021;10:100090.

    CAS  Google Scholar 

  15. Sönmez O, Turan V, Kaya C. The effects of sulfur, cattle, and poultry manure addition on soil phosphorus. Turk J Agric For. 2016;40(4):536–41.

    Article  CAS  Google Scholar 

  16. Fatma S, Hameed A, Noman M, Ahmed T, Shahid M, Tariq M, Sohail I, Tabassum R. Lignocellulosic biomass: a sustainable bioenergy source for the future. Protein Pept Lett. 2018;25(2):148–63.

    Article  CAS  PubMed  Google Scholar 

  17. Ge X, Chang C, Zhang L, Cui S, Luo X, Hu S, Qin Y, Li Y. Conversion of lignocellulosic biomass into platform chemicals for biobased polyurethane application. Adv Bioenergy. 2018;3:161–213.

    Article  CAS  Google Scholar 

  18. Zhou ZY, Liu DH, Zhao XB. Conversion of lignocellulose to biofuels and chemicals via sugar platform: an updated review on chemistry and mechanisms of acid hydrolysis of lignocellulose. Renew Sust Energy Rev. 2021;146:111169.

    Article  CAS  Google Scholar 

  19. Yang B, Wyman CE. Pretreatment: the key to unlocking low-cost cellulosic ethanol. Biofuels Bioprod Bioref. 2008;2(1):26–40.

    Article  CAS  Google Scholar 

  20. Shafiei M, Kumar R, Karimi K. Pretreatment of lignocellulosic biomass. In: Karimi K, editor. Lignocellulose-based bioproducts. Cham: Springer; 2015. p. 85–154.

    Google Scholar 

  21. Medina JDC, Magalhaes AI Jr. Ethanol production, current facts, future scenarios, and techno-economic assessment of different biorefinery configurations. Bioethanol Technologies. 2021;23:1–14.

    Google Scholar 

  22. Cunha JT, Soares PO, Baptista SL, Costa CE, Domingues L. Engineered Saccharomyces cerevisiae for lignocellulosic valorization: a review and perspectives on bioethanol production. Bioengineered. 2020;11(1):883–903.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Palmqvist E, Hahnhägerdal B. Fermentation of lignocellulosic hydrolysates. I: inhibition and detoxification. Bioresour Technol. 2000;74(1):17–24.

    Article  CAS  Google Scholar 

  24. Palmqvist E, Hahnhägerdal B. Fermentation of lignocellulosic hydrolysates. II: inhibitors and mechanisms of inhibition. Bioresour Technol. 2000;74(1):25–33.

    Article  CAS  Google Scholar 

  25. Brandt BA, Jansen T, Grgens JF, Zyl WH. Overcoming lignocellulose-derived microbial inhibitors: advancing the Saccharomyces cerevisiae resistance toolbox. Biofuel Bioprod Biorefin. 2019;13(6):1520–36.

    Article  CAS  Google Scholar 

  26. Brandt BA, García-Aparicio MDP, Görgens JF, van Zyl WH. Rational engineering of Saccharomyces cerevisiae towards improved tolerance to multiple inhibitors in lignocellulose fermentations. Biotechnol Biofuels. 2021;14:173.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Li B, Wang L, Wu YJ, Xia ZY, Yang BX, Tang YQ. Improving acetic acid and furfural resistance of xylose-fermenting Saccharomyces cerevisiae strains by regulating novel transcription factors revealed via comparative transcriptomic analysis. Appl Environ Microbiol. 2021;87(10):e00158-e221.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Domínguez E, Río PG, Romaní A, Garrote G, Domingues L. Hemicellulosic bioethanol production from fast-growing paulownia biomass. Processes. 2021;9(1):173.

    Article  CAS  Google Scholar 

  29. Ko JK, Enkh-Amgalan T, Gong G, Um Y, Lee SM. Improved bioconversion of lignocellulosic biomass by Saccharomyces cerevisiae engineered for tolerance to acetic acid. GCB Bioenergy. 2020;12(1):90–100.

    Article  CAS  Google Scholar 

  30. Sandoval-Nuñez D, Arellano-Plaza M, Gschaedler A, Arrizon J, Amaya-Delgado L. A comparative study of lignocellulosic ethanol productivities by Kluyveromyces marxianus and Saccharomyces cerevisiae. Clean Techn Environ Policy. 2018;20:1491–9.

    Article  CAS  Google Scholar 

  31. Ko JK, Um Y, Woo HM, Kim KH, Lee SM. Ethanol production from lignocellulosic hydrolysates using engineered Saccharomyces cerevisiae harboring xylose isomerase-based pathway. Bioresour Technol. 2016;209:290–6.

    Article  CAS  PubMed  Google Scholar 

  32. Romaní A, Pereira F, Johansson B, Domingues L. Metabolic engineering of Saccharomyces cerevisiae ethanol strains PE-2 and CAT-1 for efficient lignocellulosic fermentation. Bioresour Technol. 2015;179:150–8.

    Article  PubMed  CAS  Google Scholar 

  33. Martı́n C, Galbe M, Wahlbom CF, Hahn-Hägerdal B, Jönsson LJ. Ethanol production from enzymatic hydrolysates of sugarcane bagasse using recombinant xylose-utilising Saccharomyces cerevisiae. Enzyme Microb Technol. 2002;31(3):274–82.

    Article  Google Scholar 

  34. Qin L, Dong S, Yu J, Ning X, Xu K, Zhang SJ, Xu L, Li BZ, Li J, Yuan YJ, Li C. Stress-driven dynamic regulation of multiple tolerance genes improves robustness and productive capacity of Saccharomyces cerevisiae in industrial lignocellulose fermentation. Metab eng. 2020;61:160–70.

    Article  CAS  PubMed  Google Scholar 

  35. Zhu L, Li P, Sun T, Kong M, Li X, Ali S, Liu W, Fan S, Qiao J, Li S, Peng L, Hed B, Jin M, Xiao W, Cao L. Overexpression of SFA1 in engineered Saccharomyces cerevisiae to increase xylose utilization and ethanol production from different lignocellulose hydrolysates. Bioresour Technol. 2020;313:123724.

    Article  CAS  PubMed  Google Scholar 

  36. Kong M, Li X, Li T, Zhao X, Jin M, Zhou X, Gu H, Mrša V, Xiao W, Cao L. Overexpressing CCW12 in Saccharomyces cerevisiae enables highly efficient ethanol production from lignocellulose hydrolysates. Bioresour Technol. 2021;337:125487.

    Article  CAS  PubMed  Google Scholar 

  37. Caspeta L, Castillo T, Nielsen J. Modifying yeast tolerance to inhibitory conditions of ethanol production processes. Front Bioeng Biotechnol. 2015;3:184.

    Article  PubMed  PubMed Central  Google Scholar 

  38. Oh EJ, Wei N, Kwak S, Kim H, Jin YS. Overexpression of RCK1 improves acetic acid tolerance in Saccharomyces cerevisiae. J Biotechnol. 2019;292:1–4.

    Article  CAS  PubMed  Google Scholar 

  39. Cunha JT, Aguiar TQ, Romaní A, Oliveira C, Domingues L. Contribution of PRS3, RPB4 and ZWF1 to the resistance of industrial Saccharomyces cerevisiae CCUG53310 and PE-2 strains to lignocellulosic hydrolysate-derived inhibitors. Bioresour Technol. 2015;191:7–16.

    Article  CAS  PubMed  Google Scholar 

  40. Wang X, Liang Z, Hou J, Shen Y, Bao X. The absence of the transcription factor Yrr1p, identified from comparative genome profiling, increased vanillin tolerance due to enhancements of ABC transporters expressing, rRNA processing and ribosome biogenesis in Saccharomyces cerevisiae. Front Microbiol. 2017;8:367.

    PubMed  PubMed Central  Google Scholar 

  41. Li B, Xie CY, Yang BX, Gou M, Xia ZY, Sun ZY, Tang YQ. The response mechanisms of industrial Saccharomyces cerevisiae to acetic acid and formic acid during mixed glucose and xylose fermentation. Process Biochem. 2020;91:319–29.

    Article  CAS  Google Scholar 

  42. Cunha JT, Romaní A, Costa CE, Sá-Correia I, Domingues L. Molecular and physiological basis of Saccharomyces cerevisiae tolerance to adverse lignocellulose-based process conditions. Appl microbiol biot. 2019;103(1):159–75.

    Article  CAS  Google Scholar 

  43. Chen Y, Sheng J, Jiang T, Stevens J, Feng X, Wei N. Transcriptional profiling reveals molecular basis and novel genetic targets for improved resistance to multiple fermentation inhibitors in Saccharomyces cerevisiae. Biotechnol Biofuels. 2016;9(1):1–18.

    Article  CAS  Google Scholar 

  44. Zhou Z, Liu D, Zhao X. Conversion of lignocellulose to biofuels and chemicals via sugar platform: an updated review on chemistry and mechanisms of acid hydrolysis of lignocellulose. Renew Sust Energ Rev. 2021;146:111169.

    Article  CAS  Google Scholar 

  45. Robak K, Balcerek M. Review of second generation bioethanol production from residual biomass. Food Technol Biotech. 2018;56(2):174–87.

    Article  CAS  Google Scholar 

  46. Jönsson LJ, Martín C. Pretreatment of lignocellulose: formation of inhibitory by-products and strategies for minimizing their effects. Bioresour Technol. 2016;199:103–12.

    Article  PubMed  CAS  Google Scholar 

  47. Sjulander N, Kikas T. Origin, impact and control of lignocellulosic inhibitors in bioethanol production-a review. Energies. 2020;13(18):4751.

    Article  CAS  Google Scholar 

  48. Ding J, Huang X, Zhang L, Zhao N, Yang D, Zhang K. Tolerance and stress response to ethanol in the yeast Saccharomyces cerevisiae. Appl microbiol biot. 2009;85(2):253–63.

    Article  CAS  Google Scholar 

  49. Sajid M, Farooq U, Bary G, Azim MM, Zhao X. Sustainable production of levulinic acid and its derivatives for fuel additives and chemicals: progress, challenges, and prospects. Green Chem. 2021;23:9198.

    Article  CAS  Google Scholar 

  50. Wei N, Quarterman J, Kim SR, Cate JH, Jin YS. Enhanced biofuel production through coupled acetic acid and xylose consumption by engineered yeast. Nat Commun. 2013;4(1):1–8.

    Article  Google Scholar 

  51. Ndukwe JK, Aliyu GO, Onwosi CO, Chukwu KO, Ezugworie FN. Mechanisms of weak acid-induced stress tolerance in yeasts: prospects for improved bioethanol production from lignocellulosic biomass. Process Biochem. 2020;90:118–30.

    Article  CAS  Google Scholar 

  52. Stojiljkovic M, Foulquié-Moreno MR, Thevelein JM. Polygenic analysis of very high acetic acid tolerance in the yeast Saccharomyces cerevisiae reveals a complex genetic background and several new causative alleles. Biotechnol Biofuels. 2020;13(1):1–27.

    Article  CAS  Google Scholar 

  53. Antal MJ Jr, Mok WS, Richards GN. Mechanism of formation of 5-(hydroxymethyl)-2-furaldehyde from d-fructose and sucrose. Carbohydr Res. 1990;199(1):91–109.

    Article  CAS  PubMed  Google Scholar 

  54. Antal MJ Jr, Leesomboon T, Mok WS, Richards GN. Mechanism of formation of 2-furaldehyde from d-xylose. Carbohydr Res. 1991;217:71–85.

    Article  CAS  Google Scholar 

  55. Almeida JR, Modig T, Petersson A, Hähn-Hägerdal B, Lidén G, Gorwa-Grauslund MF. Increased tolerance and conversion of inhibitors in lignocellulosic hydrolysates by Saccharomyces cerevisiae. J Chem Technol Biot. 2007;82(4):340–9.

    Article  CAS  Google Scholar 

  56. Li J, Zhang M, Dowell F, Wang D. Rapid determination of acetic acid, furfural, and 5-hydroxymethylfurfural in biomass hydrolysates using near-infrared spectroscopy. ACS Omega. 2018;3(5):5355–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Wang X, Gao Q, Bao J. Transcriptional analysis of Amorphotheca resinae ZN1 on biological degradation of furfural and 5-hydroxymethylfurfural derived from lignocellulose pretreatment. Biotechnol Biofuels. 2015;8(1):136.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  58. Taherzadeh MJ, Gustafsson L, Niklasson C, Lidén G. Conversion of furfural in aerobic and anaerobic batch fermentation of glucose by Saccharomyces cerevisiae. J Biosci Bioeng. 1999;87(2):169–74.

    Article  CAS  PubMed  Google Scholar 

  59. Larsson S, Palmqvist E, Hahn-Hägerdal B, Tengborg C, Stenberg K, Zacchi G, Nilvebrant NO. The generation of fermentation inhibitors during dilute acid hydrolysis of softwood. Enzyme Microb Technol. 1999;24:151–9.

    Article  CAS  Google Scholar 

  60. Jayakody LN, Jin YS. In-depth understanding of molecular mechanisms of aldehyde toxicity to engineer robust Saccharomyces cerevisiae. Appl Microbiol Biotechnol. 2021;105:2675–92.

    Article  CAS  PubMed  Google Scholar 

  61. Liu ZL, Slininger PJ, Dien BS, Berhow MA, Kurtzman CP, Gorsich SW. Adaptive response of yeasts to furfural and 5-hydroxymethylfurfural and new chemical evidence for HMF conversion to 2,5-bis-hydroxymethylfuran. J Ind Microbiol Biot. 2004;31(8):345–52.

    Article  CAS  Google Scholar 

  62. Banerjee N, Bhatnagar R, Viswanathan L. Inhibition of glycolysis by furfural in Saccharomyces cerevisiae. Appl Microbiol Biot. 1981;11(4):226–8.

    Article  CAS  Google Scholar 

  63. Allen SA, Clark W, Mccaffery JM, Cai Z, Lanctot A, Slininger PJ, Liu ZL, Gorsich SW. Furfural induces reactive oxygen species accumulation and cellular damage in Saccharomyces cerevisiae. Biotechnol Biofuels. 2010;3(1):1–10.

    Article  CAS  Google Scholar 

  64. Fletcher E, Baetz K. Multi-faceted systems biology approaches present a cellular landscape of phenolic compound inhibition in Saccharomyces cerevisiae. Front Bioeng Biotechnol. 2020;8:539902.

    Article  PubMed  PubMed Central  Google Scholar 

  65. Gu H, Zhu Y, Peng Y, Liang X, Liu X, Shao L, Xu Y, Xu Z, Liu R, Li J. Physiological mechanism of improved tolerance of Saccharomyces cerevisiae to lignin-derived phenolic acids in lignocellulosic ethanol fermentation by short-term adaptation. Biotechnol Biofuels. 2019;12(1):1–14.

    Article  CAS  Google Scholar 

  66. López PC, Peng C, Arneborg N, Junicke H, Gernaey KV. Analysis of the response of the cell membrane of Saccharomyces cerevisiae during the detoxification of common lignocellulosic inhibitors. Sci Rep. 2021;11(1):6853.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  67. Nguyen TTM, Iwaki A, Ohya Y, Izawa S. Vanillin causes the activation of Yap1 and mitochondrial fragmentation in Saccharomyces cerevisiae. J Biosci Bioeng. 2014;117(1):33–8.

    Article  CAS  PubMed  Google Scholar 

  68. Iwaki A, Ohnuki S, Suga Y, Izawa S, Ohya Y. Vanillin inhibits translation and induces messenger ribonucleoprotein (mRNP) granule formation in Saccharomyces cerevisiae: application and validation of high-content, image-based profiling. PLoS ONE. 2013;8:e61748.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Adeboye PT, Bettiga M, Olsson L. The chemical nature of phenolic compounds determines their toxicity and induces distinct physiological responses in Saccharomyces cerevisiae in lignocellulose hydrolysates. AMB Express. 2014;4(1):1–10.

    Article  CAS  Google Scholar 

  70. Ximenes E, Kim Y, Mosier N, Dien B, Ladisch M. Deactivation of cellulases by phenols. Enzyme Microb Technol. 2011;48(1):54–60.

    Article  CAS  PubMed  Google Scholar 

  71. Qin L, Li WC, Liu L, Zhu JQ, Li X, Li BZ, Yuan YJ. Inhibition of lignin-derived phenolic compounds to cellulase. Biotechnol Biofuels. 2016;9(1):1–10.

    Article  CAS  Google Scholar 

  72. Unrean P, Gätgens J, Klein B, Noack S, Champreda V. Elucidating cellular mechanisms of Saccharomyces cerevisiae tolerant to combined lignocellulosic-derived inhibitors using high-throughput phenotyping and multiomics analyses. FEMS Yeast Res. 2018;18(8):foy106.

    Article  CAS  Google Scholar 

  73. Modig T, Almeida JR, Gorwa-Grauslund MF, Lidén G. Variability of the response of Saccharomyces cerevisiae strains to lignocellulose hydrolysate. Biotechnol Bioeng. 2008;100(3):423–9.

    Article  CAS  PubMed  Google Scholar 

  74. Udom N, Chansongkrow P, Charoensawan V, Auesukaree C. Coordination of the cell wall integrity and high-osmolarity glycerol pathways in response to ethanol stress in Saccharomyces cerevisiae. Appl Environ Microbiol. 2019;85(15):e00551-e619.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Qi Y, Liu H, Chen X, Liu L. Engineering microbial membranes to increase stress tolerance of industrial strains. Metab Eng. 2019;53:24–34.

    Article  CAS  PubMed  Google Scholar 

  76. Guo H, Qiu Y, Wei JP, Zhang YX, Yuan YH, Yue TL. Hyperosmotic stress response of Saccharomyces cerevisiae cell wall and cell membrane. T Chin Soc Agric Mach. 2020;51(6):346–52.

    Google Scholar 

  77. Klis FM, Boorsma A, De Groot PW. Cell wall construction in Saccharomyces cerevisiae. Yeast. 2006;23(3):185–202.

    Article  CAS  PubMed  Google Scholar 

  78. Klis FM, Mol P, Hellingwerf K, Brul S. Dynamics of cell wall structure in Saccharomyces cerevisiae. FEMS Microbiol Rev. 2002;26(3):239–56.

    Article  CAS  PubMed  Google Scholar 

  79. Liu ZL, Ma M. Pathway-based signature transcriptional profiles as tolerance phenotypes for the adapted industrial yeast Saccharomyces cerevisiae resistant to furfural and HMF. Appl Microbiol Biotechnol. 2020;104(8):3473–92.

    Article  CAS  PubMed  Google Scholar 

  80. Orlean P. Architecture and biosynthesis of the Saccharomyces cerevisiae cell wall. Genetics. 2012;192:775–818.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Wang Y, Zhang S, Liu H, Zhang L, Yi C, Li H. Changes and roles of membrane compositions in the adaptation of Saccharomyces cerevisiae to ethanol. J Basic Microbiol. 2015;55:1–10.

    Article  Google Scholar 

  82. Verna J, Lodder A, Lee K, Vagts A, Ballester R. A family of genes required for maintenance of cell wall integrity and for the stress response in Saccharomyces cerevisiae. P Natl Acad Sci Usa. 1997;94(25):13804–9.

    Article  CAS  Google Scholar 

  83. Boorsma A, Nobel HD, Riet BT, Bargmann B, Brul S, Hellingwerf KJ, Klis FM. Characterization of the transcriptional response to cell wall stress in Saccharomyces cerevisiae. Yeast. 2004;21(5):413–27.

    Article  CAS  PubMed  Google Scholar 

  84. Alonso-Monge R, Real E, Wojda I, Bebelman JP, Mager WH, Siderius M. Hyperosmotic stress response and regulation of cell wall integrity in Saccharomyces cerevisiae share common functional aspects. Mol microbiol. 2001;41(3):717–30.

    Article  CAS  PubMed  Google Scholar 

  85. Maaheimo H, Fiaux J, Cakar ZP, Bailey JE, Sauer U, Szyperski T. Central carbon metabolism of Saccharomyces cerevisiae explored by biosynthetic fractional 13C labeling of common amino acids. Eur J Biochem. 2001;268(8):2464–79.

    Article  CAS  PubMed  Google Scholar 

  86. Li BZ, Yuan YJ. Transcriptome shifts in response to furfural and acetic acid in Saccharomyces cerevisiae. Appl Microbiol Biotechnol. 2010;86(6):1915–24.

    Article  CAS  PubMed  Google Scholar 

  87. Guo Z, Olsson L. Physiological response of Saccharomyces cerevisiae to weak acids present in lignocellulosic hydrolysate. FEMS Yeast Res. 2014;14(8):1234–48.

    Article  CAS  PubMed  Google Scholar 

  88. Shui W, Xiong Y, Xiao W, Qi X, Zhang Y, Lin Y, Guo Y, Zhang Z, Wang Q, Ma Y. Understanding the mechanism of thermotolerance distinct from heat shock response through proteomic analysis of industrial strains of Saccharomyces cerevisiae. Mol Cell Proteomics. 2015;14(7):1885–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Leite FC, Leite DV, Pereira LF, de Barros PW, de Morais MA, Jr. High intracellular trehalase activity prevents the storage of trehalose in the yeast Dekkera bruxellensis. Lett Appl Microbiol. 2016;63(3):210–4.

    Article  CAS  PubMed  Google Scholar 

  90. Ribeiro RA, Vitorino MV, Godinho CP, Bourbon-Melo N, Robalo TT, Fernandes F, Rodrigues MS, Sá-Correia I. Yeast adaptive response to acetic acid stress involves structural alterations and increased stiffness of the cell wall. Sci Rep. 2021;11(1):1–9.

    Article  CAS  Google Scholar 

  91. Mollapour M, Shepherd A, Piper PW. Presence of the Fps1p aquaglyceroporin channel is essential for Hog1p activation, but suppresses Slt2(Mpk1)p activation, with acetic acid stress of yeast. Microbiology. 2009;155(10):3304–11.

    Article  CAS  PubMed  Google Scholar 

  92. Mat Nanyan NSB, Takagi H. Proline homeostasis in Saccharomyces cerevisiae: How does the stress-responsive transcription factor Msn2 play a role? Front Genet. 2020;11:438.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  93. Takagi H. Proline as a stress protectant in yeast: physiological functions, metabolic regulations, and biotechnological applications. Appl Microbiol Biotechnol. 2008;81(2):211–23.

    Article  CAS  PubMed  Google Scholar 

  94. Nishimura A, Kotani T, Sasano Y, Takagi H. An antioxidative mechanism mediated by the yeast N-acetyltransferase Mpr1: oxidative stress induced arginine synthesis and its physiological role. FEMS Yeast Res. 2010;10(6):687–98.

    Article  CAS  PubMed  Google Scholar 

  95. Schroeder L, Ikui AE. Tryptophan confers resistance to SDS-associated cell membrane stress in Saccharomyces cerevisiae. PLoS ONE. 2019;14(3):e0199484.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Mira NP, Palma M, Guerreiro JF, Sá-Correia I. Genome-wide identification of Saccharomyces cerevisiae genes required for tolerance to acetic acid. Microb Cell Fact. 2010;9(1):79.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  97. Dong Y, Hu J, Fan L, Chen Q. RNA-Seq-based transcriptomic and metabolomic analysis reveal stress responses and programmed cell death induced by acetic acid in Saccharomyces cerevisiae. Sci Rep. 2017;7:42659.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Crawford RA, Pavitt GD. Translational regulation in response to stress in Saccharomyces cerevisiae. Yeast. 2019;36(1):5–21.

    Article  CAS  PubMed  Google Scholar 

  99. Rolfes RJ. Regulation of purine nucleotide biosynthesis: in yeast and beyond. Biochem Soc Trans. 2006;34:786–90.

    Article  CAS  PubMed  Google Scholar 

  100. Zhang MM, Xiong L, Tang YJ, Mehmood MA, Zhao ZK, Bai FW, Zhao XQ. Enhanced acetic acid stress tolerance and ethanol production in Saccharomyces cerevisiae by modulating expression of the de novo purine biosynthesis genes. Biotechnol Biofuels. 2019;12(1):1–13.

    Article  Google Scholar 

  101. Ishida Y, Nguyen TT, Kitajima S, Izawa S. Prioritized expression of BDH2 under bulk translational repression and its contribution to tolerance to severe vanillin stress in Saccharomyces cerevisiae. Front Microbiol. 2016;7:1059.

    Article  PubMed  PubMed Central  Google Scholar 

  102. Nguyen TTM, Kitajima S, Izawa S. Importance of glucose-6-phosphate dehydrogenase (G6PDH) for vanillin tolerance in Saccharomyces cerevisiae. J Biosci Bioeng. 2014;118:263–9.

    Article  CAS  PubMed  Google Scholar 

  103. Liang Z, Wang X, Bao X, Wei T, Hou J, Liu W, Shen Y. Newly identified genes contribute to vanillin tolerance in Saccharomyces cerevisiae. Microb Biotechnol. 2021;14(2):503–16.

    Article  CAS  PubMed  Google Scholar 

  104. Drakulic T, Temple MD, Guido R, Jarolim S, Breitenbach M, Attfield PV, Dawes IW. Involvement of oxidative stress response genes in redox homeostasis, the level of reactive oxygen species, and ageing in Saccharomyces cerevisiae. FEMS Yeast Res. 2005;5(12):1215–28.

    Article  CAS  PubMed  Google Scholar 

  105. Zhang M, Zhang K, Mehmood MA, Zhao ZK, Bai F, Zhao X. Deletion of acetate transporter gene ADY2 improved tolerance of Saccharomyces cerevisiae against multiple stresses and enhanced ethanol production in the presence of acetic acid. Bioresour Technol. 2017;245(Pt B):1461–8.

    Article  CAS  PubMed  Google Scholar 

  106. Zhang MM, Zhao XQ, Cheng C, Bai FW. Improved growth and ethanol fermentation of Saccharomyces cerevisiae in the presence of acetic acid by overexpression of SET5 and PPR1. Biotechnol J. 2015;10(12):1903–11.

    Article  CAS  PubMed  Google Scholar 

  107. Wu X, Zhang L, Jin X, Fang Y, Zhang K, Qi L, Zheng D. Deletion of JJJ1 improves acetic acid tolerance and bioethanol fermentation performance of Saccharomyces cerevisiae strains. Biotechnol Lett. 2016;38(7):1097–106.

    Article  CAS  PubMed  Google Scholar 

  108. Morano KA, Grant CM, Moye-Rowley WS. The response to heat shock and oxidative stress in Saccharomyces cerevisiae. Genetics. 2012;190(4):1157–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Xu K, Gao L, Hassan JU, Zhao Z, Li C, Huo YX, Liu G. Improving the thermo-tolerance of yeast base on the antioxidant defense system. Chem Eng Sci. 2018;175:335–42.

    Article  CAS  Google Scholar 

  110. Mollapour M, Piper PW. Hog1 mitogen-activated protein kinase phosphorylation targets the yeast Fps1 aquaglyceroporin for endocytosis, thereby rendering cells resistant to acetic acid. Mol cell boil. 2007;27(18):6446–56.

    Article  CAS  Google Scholar 

  111. Hohmann S. Osmotic stress signaling and osmoadaptation in yeasts. Microbiol Mol Biol Rev. 2002;66(2):300–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Sundström L, Larsson S, Jönsson LJ. Identification of Saccharomyces cerevisiae genes involved in the resistance to phenolic fermentation inhibitors. Appl Biochem Biotechnol. 2010;161(1):106–15.

    Article  PubMed  CAS  Google Scholar 

  113. Tenreiro S, Rosa PC, Viegas CA, Sá-Correia I. Expression of the AZR1 gene (ORF YGR224w), encoding a plasma membrane transporter of the major facilitator superfamily, is required for adaptation to acetic acid and resistance to azoles in Saccharomyces cerevisiae. Yeast. 2000;16(16):1469–81.

    Article  CAS  PubMed  Google Scholar 

  114. Lee Y, Nasution O, Lee YM, Kim E, Choi W, Kim W. Overexpression of PMA1 enhances tolerance to various types of stress and constitutively activates the SAPK pathways in Saccharomyces cerevisiae. Appl microbiol biot. 2017;101(1):229–39.

    Article  CAS  Google Scholar 

  115. Lei J, Yu S, Xu L, Peng B, Xiao Y, Bao X. Enhanced resistance of Saccharomyces cerevisiae to vanillin by expression of lacA from Trametes sp. ah28–2. Bioresour Technol. 2011;102(17):8105–9.

    Article  CAS  Google Scholar 

  116. Ding J, Holzwarth G, Penner MH, Patton-Vogt J, Bakalinsky AT. Overexpression of acetyl-CoA synthetase in Saccharomyces cerevisiae increases acetic acid tolerance. FEMS Microbiol Lett. 2015;362(3):1–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Hasunuma T, Ismail KSK, Nambu Y, Kondo A. Co-expression of TAL1 and ADH1 in recombinant xylose-fermenting Saccharomyces cerevisiae improves ethanol production from lignocellulosic hydrolysates in the presence of furfural. J Biosci Bioeng. 2014;117(2):165–9.

    Article  CAS  PubMed  Google Scholar 

  118. Almeida JRM, Roder A, Modig T, Laadan B, Liden G, Gorwa-Grauslund MF. NADH-vs NADPH-coupled reduction of 5-hydroxymethyl furfural (HMF) and its implications on product distribution in Saccharomyces cerevisiae. Appl Microbiol Biotechnol. 2008;78(6):939–45.

    Article  CAS  PubMed  Google Scholar 

  119. Petersson A, Almeida JR, Modig T, Karhumaa K, Hahn-Hagerdal B, Gorwa-Grauslund MF, Liden G. A 5-hydroxymethyl furfural reducing enzyme encoded by the Saccharomyces cerevisiae ADH6 gene conveys HMF tolerance. Yeast. 2006;23(6):455–64.

    Article  CAS  PubMed  Google Scholar 

  120. Ma M, Liu ZL. Comparative transcriptome profiling analyses during the lag phase uncover YAP1, PDR1, PDR3, RPN4, and HSF1 as key regulatory genes in genomic adaptation to the lignocellulose derived inhibitor HMF for Saccharomyces cerevisiae. BMC Genom. 2010;11(1):1–19.

    Article  CAS  Google Scholar 

  121. Park SE, Koo HM, Park YK, Park SM, Park JC, Lee OK, Park YC, Seo JH. Expression of aldehyde dehydrogenase 6 reduces inhibitory effect of furan derivatives on cell growth and ethanol production in Saccharomyces cerevisiae. Bioresour Technol. 2011;102(10):6033–8.

    Article  CAS  PubMed  Google Scholar 

  122. Heer D, Heine D, Sauer U. Resistance of Saccharomyces cerevisiae to high concentrations of furfural is based on nadph-dependent reduction by at least two oxireductases. Appl Environ Microbiol. 2009;75(24):7631–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Liu CG, Li K, Li KY, Sakdaronnarong C, Mehmood MA, Zhao XQ, Bai FW. Intracellular redox perturbation in Saccharomyces cerevisiae improved furfural tolerance and enhanced cellulosic bioethanol production. Front Bioeng Biotech. 2020;8:615.

    Article  Google Scholar 

  124. Larsson S, Nilvebrant NO, Jonsson LJ. Effect of overexpression of Saccharomyces cerevisiae Pad1p on the resistance to phenylacrylic acids and lignocellulose hydrolysates under aerobic and oxygen-limited conditions. Appl Microbiol Biotechnol. 2001;57:167–74.

    Article  CAS  PubMed  Google Scholar 

  125. Li YC, Gou ZX, Liu ZS, Tang YQ, Akamatsu T, Kida K. Synergistic effects of TAL1 over-expression and PHO13 deletion on the weak acid inhibition of xylose fermentation by industrial Saccharomyces cerevisiae strain. Biotechnol Lett. 2014;36(10):2011–21.

    Article  CAS  PubMed  Google Scholar 

  126. Fujitomi K, Sanda T, Hasunuma T, Kondo A. Deletion of the PHO13 gene in Saccharomyces cerevisiae improves ethanol production from lignocellulosic hydrolysate in the presence of acetic and formic acids, and furfural. Bioresour Technol. 2012;111:161–6.

    Article  CAS  PubMed  Google Scholar 

  127. Chen Y, Stabryla L, Wei N. Improved acetic acid resistance in Saccharomyces cerevisiae by overexpression of the WHI2 gene identified through inverse metabolic engineering. Appl Environ Microbiol. 2016;82(7):2156–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Gorsich SW, Dien BS, Nichols NN, Slininger PJ, Liu ZL, Skory CD. Tolerance to furfural-induced stress is associated with pentose phosphate pathway genes ZWF1, GND1, RPE1, and TKL1 in Saccharomyces cerevisiae. Appl Microbiol Biotechnol. 2006;71:339–49.

    Article  CAS  PubMed  Google Scholar 

  129. Ask M, Mapelli V, Höck H, Olsson L, Bettiga M. Engineering glutathione biosynthesis of Saccharomyces cerevisiae increases robustness to inhibitors in pretreated lignocellulosic materials. Microb Cell Fact. 2013;12(1):1–10.

    Article  CAS  Google Scholar 

  130. Swinnen S, Henriques SF, Shrestha R, Ho PW, Sá-Correia I, Nevoigt E. Improvement of yeast tolerance to acetic acid through Haa1 transcription factor engineering: towards the underlying mechanisms. Microb Cell Fact. 2017;16(7):1–15.

    Google Scholar 

  131. Fernandes AR, Mira NP, Vargas RC, Canelhas I, Sá-Correia I. Saccharomyces cerevisiae adaptation to weak acids involves the transcription factor Haa1p and Haa1p-regulated genes. Biochem Biophys Res Commun. 2005;337(1):95–103.

    Article  CAS  PubMed  Google Scholar 

  132. Tanaka K, Ishii Y, Ogawa J, Shima J. Enhancement of acetic acid tolerance in Saccharomyces cerevisiae by overexpression of the HAA1 gene, encoding a transcriptional activator. Appl Environ Microb. 2012;78(22):8161–3.

    Article  CAS  Google Scholar 

  133. Sasano Y, Watanabe D, Ukibe K, Inai T, Ohtsu I, Shimoi H, Takagi H. Overexpression of the yeast transcription activator Msn2 confers furfural resistance and increases the initial fermentation rate in ethanol production. J Biosci Bioeng. 2012;113(4):451–5.

    Article  CAS  PubMed  Google Scholar 

  134. Watanabe M, Watanabe D, Akao T, Shimoi H. Overexpression of MSN2 in a sake yeast strain promotes ethanol tolerance and increases ethanol production in sake brewing. J Biosci Bioeng. 2009;107(5):516–8.

    Article  CAS  PubMed  Google Scholar 

  135. Sadeh A, Movshovich N, Volokh M, Gheber L, Aharoni A. Fine-tuning of the Msn2/4-mediated yeast stress responses as revealed by systematic deletion of Msn2/4 partners. Mol Biol Cell. 2011;22(17):3127–38.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Morgan BA, Banks GR, Toone WM, Raitt D, Kuge S, Johnston LH. The Skn7 response regulator controls gene expression in the oxidative stress response of the budding yeast Saccharomyces cerevisiae. EMBO J. 1997;16(5):1035–44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Mulford KE, Fassler JS. Association of the Skn7 and Yap1 transcription factors in the Saccharomyces cerevisiae oxidative stress response. Eukaryot Cell. 2011;10(6):761–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. He XJ, Mulford KE, Fassler JS. Oxidative stress function of the Saccharomyces cerevisiae Skn7 receiver domain. Eukaryot Cell. 2009;8(5):768–78.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Wu G, Xu Z, Jönsson LJ. Profiling of Saccharomyces cerevisiae transcription factors for engineering the resistance of yeast to lignocellulose-derived inhibitors in biomass conversion. Microb Cell Fact. 2017;16(1):199.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  140. Kim D, Hahn JS. Roles of the yap1 transcription factor and antioxidants in Saccharomyces cerevisiae’s tolerance to furfural and 5-hydroxymethylfurfural, which function as thiol-reactive electrophiles generating oxidative stress. Appl Environ Microbiol. 2013;79(16):5069–77.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. De Witt RN, Kroukamp H, Volschenk H. Proteome response of two natural strains of Saccharomyces cerevisiae with divergent lignocellulosic inhibitor stress tolerance. FEMS Yeast Res. 2019;19(1):foy116.

    Google Scholar 

  142. Xu K, Qin L, Bai W, Wang X, Li F, Ren S, Gao X, Chen B, Tong Y, Li J, Li BZ, Yuan YJ, Li C. Multilevel defense system (MDS) relieves multiple stresses for economically boosting ethanol production of industrial Saccharomyces cerevisiae. ACS Energy Lett. 2020;5(2):572–82.

    Article  CAS  Google Scholar 

  143. Mertens JA, Skory CD, Nichols NN, Hector RE. Impact of stress-response related transcription factor overexpression on lignocellulosic inhibitor tolerance of Saccharomyces cerevisiae environmental isolates. Biotechnol Prog. 2021;37(2):e3094.

    Article  CAS  PubMed  Google Scholar 

  144. Mira NP, Becker JD, Sá-Correia I. Genomic expression program involving the Haa1p-regulon in Saccharomyces cerevisiae response to acetic acid. OMICS. 2010;14(5):587–601.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Abbott DA, Suir E, Van Maris AJA, Pronk JT. Physiological and transcriptional responses to high concentrations of lactic acid in anaerobic chemostat cultures of Saccharomyces cerevisiae. Appl Environ Microbiol. 2008;74(18):5769–5768.

    Article  CAS  Google Scholar 

  146. Sugiyama M, Akase SP, Nakanishi R, Horie H, Kaneko Y, Harashima S. Nuclear localization of Haa1, which is linked to its phosphorylation status, mediates lactic acid tolerance in Saccharomyces cerevisiae. Appl Environ Microb. 2014;80(11):3488–95.

    Article  CAS  Google Scholar 

  147. Collins ME, Black JJ, Liu Z. Casein kinase I isoform Hrr25 is a negative regulator of Haa1 in the weak acid stress response pathway in Saccharomyces cerevisiae. Appl Environ Microb. 2017;83(13):e00672-e717.

    Article  CAS  Google Scholar 

  148. Kim MS, Cho KH, Park KH, Jang J, Hahn JS. Activation of Haa1 and War1 transcription factors by differential binding of weak acid anions in Saccharomyces cerevisiae. Nucleic Acids Res. 2019;47(3):1211–24.

    Article  CAS  PubMed  Google Scholar 

  149. Mira NP, Henriques SF, Keller G, Teixeira MC, Matos RG, Arraiano CM, Winge DR, Sa-Correia I. Identification of a DNA-binding site for the transcription factor Haa1, required for Saccharomyces cerevisiae response to acetic acid stress. Nucleic Acids Res. 2011;39(16):6896–907.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Keller G, Ray E, Brown PO, Winge DR. Haa1, a protein homologous to the copper-regulated transcription factor Ace1, is a novel transcriptional activator. J Biol Chem. 2001;276(42):38697–702.

    Article  CAS  PubMed  Google Scholar 

  151. Sakihama Y, Hasunuma T, Kondo A. Improved ethanol production from xylose in the presence of acetic acid by the overexpression of the HAA1 gene in Saccharomyces cerevisiae. J Biosci Bioeng. 2015;119(3):297–302.

    Article  CAS  PubMed  Google Scholar 

  152. Gasch AP, Spellman PT, Kao CM, Carmel-Harel O, Eisen MB, Storz G, Botstein D, Brown PO. Genomic expression programs in the response of yeast cells to environmental changes. Mol Biol Cell. 2000;11(12):4241–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Beck T, Hall MN. The TOR signalling pathway controls nuclear localization of nutrient-regulated transcription factors. Nature. 1999;402(6762):689–92.

    Article  CAS  PubMed  Google Scholar 

  154. Görner W, Durchschlag E, Martinez-Pastor MT, Estruch F, Ammerer G, Hamilton B, Ruis H, Schüller C. Nuclear localization of the C2H2 zinc finger protein Msn2p is regulated by stress and protein kinase A activity. Genes Dev. 1998;12(4):586–97.

    Article  PubMed  PubMed Central  Google Scholar 

  155. Berry DB, Gasch AP. Stress-activated genomic expression changes serve a preparative role for impending stress in yeast. Mol Biol Cell. 2008;19(11):4580–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Moskvina E, Schüller C, Maurer CTC, Mager WH, Ruis H. A search in the genome of Saccharomyces cerevisiae for genes regulated via stress response elements. Yeast. 1998;14(11):1041–50.

    Article  CAS  PubMed  Google Scholar 

  157. Görner W, Durchschlag E, Wolf J, Brown EL, Ammerer G, Ruis H, Schüller C. Acute glucose starvation activates the nuclear localization signal of a stress-specific yeast transcription factor. Embo J. 2002;21(1–2):135–44.

    Article  PubMed  PubMed Central  Google Scholar 

  158. Boy-Marcotte E, Garmendia C, Garreau H, Lallet S, Mallet L, Jacquet M. The transcriptional activation region of Msn2p, in Saccharomyces cerevisiae, is regulated by stress but is insensitive to the cAMP signalling pathway. Mol Genet Genomics. 2006;275(3):277–87.

    Article  CAS  PubMed  Google Scholar 

  159. Eastmond DL, Nelson HCM. Genome-wide analysis reveals new roles for the activation domains of the Saccharomyces cerevisiae heat shock transcription factor (Hsf1) during the transient heat shock response. J Biol Chem. 2006;281:32909–21.

    Article  CAS  PubMed  Google Scholar 

  160. Causton HC, Ren B, Koh SS, Harbison CT, Kanin E, Jennings EG, Lee TI, True HL, Lander ES, Young RA. Remodeling of yeast genome expression in response to environmental changes. Mol Biol Cell. 2001;12(2):323–37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Schnell N, Krems B, Entian KD. The PAR1 (YAP1/SNQ3) gene of Saccharomyces cerevisiae, ac-jun homologue, is involved in oxygen metabolism. Curr genet. 1992;21(4):269–73.

    Article  CAS  PubMed  Google Scholar 

  162. Kuge S, Jones N. YAP1 dependent activation of TRX2 is essential for the response of Saccharomyces cerevisiae to oxidative stress by hydroperoxides. Embo J. 1994;13(3):655–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Moye-Rowley WS, Harshman KD, Parker CS. Yeast YAP1 encodes a novel form of the jun family of transcriptional activator proteins. Genes Dev. 1989;3(3):283–92.

    Article  CAS  PubMed  Google Scholar 

  164. Rodriguez-Colman MJ, Reverter-Branchat G, Sorolla MA, Tamarit J, Ros J, Cabiscol E. The forkhead transcription factor Hcm1 promotes mitochondrial biogenesis and stress resistance in yeast. J Biol Chem. 2010;285(47):37092–101.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  165. Shen MJ, Wu Y, Yang K, Li Y, Xu H, Zhang H, Li BZ, Li X, Xiao WH, Zhou X, Mitchell LA, Bader JS, Yuan YJ, Boeke JD. Heterozygous diploid and interspecies SCRaMbLEing. Nat commun. 2018;9(1):1–8.

    Article  CAS  Google Scholar 

  166. Shen Y, Stracquadanio G, Wang Y, Yang K, Mitchell LA, Xue Y, Cai Y, Chen T, Dymond JS, Kang K, Gong J, Zeng X, Zhang Y, Li Y, Feng Q, Xu X, Wang J, Wang J, Yang H, Boeke JD, Bader JS. SCRaMbLE generates designed combinatorial stochastic diversity in synthetic chromosomes. Genome Res. 2016;26(1):36–49.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  167. Dymond J, Boeke J. The Saccharomyces cerevisiae SCRaMbLE system and genome minimization. Bioengineered. 2012;3(3):170–3.

    Article  Google Scholar 

  168. Alper H, Moxley J, Nevoigt E, Fink GR. Stephanopoulos G. Engineering yeast transcription machinery for improved ethanol tolerance and production. Science. 2006;314(5805):1565–8.

    Article  CAS  PubMed  Google Scholar 

  169. Si T, Luo Y, Bao Z, Zhao H. RNAi-assisted genome evolution in Saccharomyces cerevisiae for complex phenotype engineering. ACS Synth Biol. 2015;4(3):283–91.

    Article  CAS  PubMed  Google Scholar 

  170. Si T, Chao R, Min Y, Wu Y, Ren W, Zhao H. Automated multiplex genome-scale engineering in yeast. Nat commun. 2017;8(1):1–12.

    Article  Google Scholar 

  171. Luo SH, Huang YW, Chen A, Wang JL. Domestication and screening of Saccharomyces cerevisiae strain resistant to inhibitors in lignocellulosic hydrolysates by acclimatizing inhibitory. Appl Mech Mater. 2013;448:1581–6.

    Article  Google Scholar 

  172. Almario MP, Reyes LH, Kao KC. Evolutionary engineering of Saccharomyces cerevisiae for enhanced tolerance to hydrolysates of lignocellulosic biomass. Biotechnol bioeng. 2013;110(10):2616–23.

    Article  CAS  PubMed  Google Scholar 

  173. Ottenheim C, Nawrath M, Wu JC. Microbial mutagenesis by atmospheric and room-temperature plasma (ARTP): the latest development. Bioresour Bioprocess. 2018;5(1):1–14.

    Article  Google Scholar 

  174. Xie CY, Yang BX, Wu YJ, Xia ZY, Gou M, Sun ZY, Tang YQ. Construction of industrial xylose-fermenting Saccharomyces cerevisiae strains through combined approaches. Process Biochem. 2020;96:80–9.

    Article  CAS  Google Scholar 

  175. Casal M, Cardoso H, Leao C. Mechanisms regulating the transport of acetic acid in Saccharomyces cerevisiae. Microbiology. 1996;142(6):1385–90.

    Article  CAS  PubMed  Google Scholar 

  176. Teymennet-Ramírez KV, Martínez-Morales F, Muñoz-Garay C, Bertrand B, Morales-Guzmán D, Trejo-Hernández MR. Laccase treatment of phenolic compounds for bioethanol production and the impact of these compounds on yeast physiology. Biocatal Biotransfor. 2020;40(1):38–49.

    Article  CAS  Google Scholar 

Download references

Funding

This work was supported by the National Key R&D Program of China (2018YFA0902200) and National Natural Science Foundation of China (No. 21878176).

Author information

Authors and Affiliations

Authors

Contributions

BL wrote the draft of the manuscript. NL participated in writing a part of the manuscript; XZ edited and revised the manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Xuebing Zhao.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, B., Liu, N. & Zhao, X. Response mechanisms of Saccharomyces cerevisiae to the stress factors present in lignocellulose hydrolysate and strategies for constructing robust strains. Biotechnol Biofuels 15, 28 (2022). https://doi.org/10.1186/s13068-022-02127-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13068-022-02127-9

Keywords