Abstract
Fungi exhibit remarkable morphological plasticity, which allows them to undergo reversible transitions between distinct cellular states in response to changes in their environment. This phenomenon, termed fungal morphogenesis, is critical for fungi to survive and colonize diverse ecological niches and establish infections in a variety of hosts. Despite significant advancements in the field with respect to understanding the gene regulatory networks that control these transitions, the metabolic determinants of fungal morphogenesis remain poorly characterized. In this study, we uncover a previously uncharacterized, conserved dependency between central carbon metabolism and de novo biosynthesis of sulfur-containing amino acids that is critical for fungal morphogenesis, in two key fungal species. Using a multidisciplinary approach, we demonstrate that glycolytic flux is crucial to drive fungal morphogenesis in a cAMP-independent manner and perturbation of this pathway leads to a significant downregulation in the expression of genes involved in de novo biosynthesis of sulfur-containing amino acids. Remarkably, exogenous supplementation of sulfur-containing amino acids robustly rescues the morphogenesis defect induced by the perturbation of glycolysis in both Saccharomyces cerevisiae and Candida albicans, underscoring the pivotal role of de novo biosynthesis of sulfur-containing amino acid as a downstream effector of morphogenesis. Furthermore, a C. albicans mutant lacking the glycolytic enzyme, phosphofructokinase (Pfk1) exhibited significantly reduced survival within murine macrophages and attenuated virulence in a murine model of systemic candidiasis. Overall, our work elucidates a previously uncharacterized coupling between glycolysis and sulfur metabolism that is critical for driving fungal morphogenesis, contributing to our understanding of this conserved phenomenon.
Introduction
In response to changes in their environment, fungi exhibit remarkable morphological plasticity, allowing them to reversibly transition between specialized cell states. This inherent ability to rapidly adapt to the changes in their environment is essential for their survival and allows them to colonize diverse ecological niches including various host organisms (Gow et al., 2002; Lin et al., 2015; Naranjo-Ortiz and Gabaldón, 2019). This conserved phenomenon, of fungal morphogenesis, encompasses a plethora of morphological forms including pseudohyphae and hyphae to name a few (Lin et al., 2015). Importantly, these morphological transitions in various pathogenic fungal species significantly influence their interaction with the respective host, impacting their overall ability to cause persistent infections (Min et al., 2020; Wang and Lin, 2012). Fungal morphogenesis is conserved across a plethora of fungal species and has profound implications with respect to fungal ecology, anti-fungal drug resistance and pathogenicity across a spectrum of hosts (Christensen, 1989; Puumala et al., 2024; Sexton and Howlett, 2006). Consequently, it has been a central focus of intense scientific investigation over the last several years (Lin et al., 2015; Riquelme et al., 2018). During dimorphic transitions in fungi, changes in cell shape is accompanied with changes in cell polarity, cell size and cell-cell adhesion and these changes are regulated by two major signalling pathways: the cAMP-PKA pathway and MAP kinase (MAPK) pathways. Both pathways are crucial for morphogenesis in Saccharomyces cerevisiae and Candida albicans. The cAMP-PKA pathway acts as a key nutrient sensor; specifically, the presence of glucose activates it to promote pseudohyphal differentiation in S. cerevisiae (Xue et al., 1998). Similarly, this pathway drives the essential yeast-to-hyphal transition in C. albicans, often via the regulator Efg1 (Maidan et al., 2005). Concurrently, the MAPK pathway responds to various environmental stresses and plays a crucial role in regulating fungal morphogenesis. For example, in S. cerevisiae, the Kss1 pathway is responsible for driving filamentous growth (Lorenz and Heitman, 1998). Similarly, in C. albicans, the Cek1 pathway controls hyphal development and biofilm formation (Csank et al., 1998). The cAMP-PKA and MAPK pathways function in a parallel manner, engaging in significant crosstalk to integrate diverse environmental signals and coordinate a precise morphological response (Biswas et al., 2007; Kumar, 2021). While significant progress has been made in elucidating the intricate gene regulatory networks that govern these morphological transitions (e.g., Genetic networks that orchestrate yeast to pseudohyphal differentiation in S. cerevisiae and yeast to hyphal differentiation in C. albicans have been well characterized (Nantel et al., 2002; Ryan et al., 2012)), a critical knowledge gap persists regarding the underlying metabolic determinants of fungal morphogenesis. Despite the recognized influence of nutrient availability on fungal morphogenesis (Broach, 2012; Fleck et al., 2011; Kumar, 2021), our understanding of the specific metabolic networks and their regulatory influence on morphogenetic switching in fungi remain incompletely characterized.
In this study, we employed S. cerevisiae and the human fungal pathogen, C. albicans to dissect the metabolic underpinnings of fungal morphogenesis, under nitrogen-limiting conditions, a known trigger for morphogenetic switching in these fungi (Csank et al., 1998; Gimeno et al., 1992). Our findings reveal a critical role for central carbon metabolism, particularly glycolysis, in facilitating the yeast-to-pseudohyphal transition in S. cerevisiae in a cAMP-PKA independent manner and the yeast-to-hyphal transition in C. albicans, both of which have been shown to be crucial for their nutrient foraging ability in diverse niches and the ability of C. albicans to successfully infect a host (Palková and Váchová, 2006; Thompson et al., 2011). Comparative transcriptomic analysis identified a conserved metabolic network crucial for these morphogenetic switching events under nitrogen-limiting conditions. Importantly, we observed that perturbations in glycolytic flux using pharmacological and genetic means, negatively impact the expression of genes involved in sulfur metabolism. Furthermore, our functional studies demonstrated that supplementation with sulfur-containing amino acids (cysteine or methionine), effectively rescued the morphogenetic defects resulting from glycolytic impairment, establishing a novel metabolic axis linking glycolysis and sulfur metabolism in the regulation of fungal morphogenesis. Finally, we investigated the implications of this glycolysis-dependent sulfur metabolism on the virulence of C. albicans. Our data demonstrates that a C. albicans mutant lacking genes that encode for a crucial glycolytic enzyme, phosphofructokinase (Pfk1), exhibits reduced survival within macrophages and severely attenuated virulence in an in vivo murine model of systemic candidiasis which is rescued in response to sulfur supplementation to the host.
Collectively, our research, for the first time, identifies a previously uncharacterized metabolic axis linking glycolysis and sulfur metabolism under conditions that induce fungal morphogenesis and underscores the importance of this metabolic network in regulating fungal morphogenesis across two key fungal species and the ability of the human fungal pathogen, C. albicans to establish infection in a host. These findings address a fundamental knowledge gap in the field and contributes significantly to our understanding of a broadly conserved phenomenon that exists across a plethora of fungal species.
Results
Glycolysis is Critical for Pseudohyphal Differentiation in a cAMP-PKA Independent Manner in S. cerevisiae
Nitrogen limitation is essential for yeast to pseudohyphal transition in S. cerevisiae (Gimeno et al., 1992; Pan et al., 2000). However, the influence of carbon sources on this phenomenon is not completely understood. It has been observed that fermetable carbon sources (sucrose, maltose, glucose, etc.) are important for inducing pseudohyphal differntiation under nitrogen-limiting conditions (Van De Velde and Thevelein, 2008). Conversely, when these were replaced with non-fermentable carbon sources, there was a significant reduction in pseudohyphal differentiation (Strudwick et al., 2010). This indicates that the presence of fermentable carbon sources is critical for pseudohyphal differentiation in S. cerevisiae. Crabtree positive organisms like S. cerevisiae metabolize glucose primarily via glycolysis to meet their growth and energy demands (Barford and Hall, 1979; De Deken, 1966). In order to determine whether the ability of S. cerevisiae to metabolize glucose via glycolysis is critical for pseudohyphal differentiation under nitrogen-limiting conditions, we spotted S. cerevisiae wild-type ∑1278b on nitrogen-limiting media containing 2% glucose (SLAD) with and without sub-inhibitory concentrations of 2-Deoxy-D-Glucose (2DG) and sodium citrate (NaCi), which are well characterized glycolysis inhibitors (Cramer and Woodward, 1952; Evans and Ratledge, 1985; Yoshino and Murakami, 1982) and monitored pseudohyphal differentiation (Fig. 1A). Our data demonstrates that sub-inhibitory concentrations of 2DG and NaCi significantly attenuated the ability of S. cerevisiae to undergo pseudohyphal differentiation (Fig. 1B). We then isolated cells from these colonies and quantified the percentage of pseudohyphal cells (cells with a length/width ratio of 2 or more (Schröder et al., 2000)) in these colonies. Addition of 2DG or NaCi resulted in a significant reduction in pseudohyphal cells (Fig. 1C) without compromising the overall growth of S. cerevisiae wild-type ∑1278b (Fig. 1G).

Glycolysis is Critical for Pseudohyphal Differentiation in a cAMP-PKA Independent Manner in S. cerevisiae.
(A) Schematic overview of glycolysis and its inhibition by glycolysis inhibitors (2-Deoxy-D-Glucose (2DG) and sodium citrate (NaCi)). (B) Wild-type ∑1278b or CEN.PK were spotted on nitrogen-limiting medium containing 2% glucose (SLAD) with and without sub-inhibitory concentrations of glycolysis inhibitors (2-Deoxy-D-Glucose (2DG) and sodium citrate (NaCi)), incubated for 10 days at 30 °C. Whole colony imaging was done using Olympus MVX10 stereo microscope and single cell imaging was done using Zeiss Apotome microscope. Scale bar represents 1mm for whole colony images and 5µm for single cell images. (C) Cells from colonies were isolated and single cells were imaged using Zeiss Apotome microscope and percentage of pseudohyphal cells were calculated using ImageJ. More than 500 cells were counted for each condition. Statistical analysis was done using one-way ANOVA test, ***(P<0.001). Error bars represent SEM. (D) Schematic overview of glycolysis showing targeted gene deletions. (E) Pertinent knockout strains which lack genes encoding for enzymes involved in glycolysis such as pfk1 and adh1 were spotted on SLAD along with wild-type ∑1278b or CEN.PK. Whole colony imaging was done using Olympus MVX10 stereo microscope and single cell imaging was done using Zeiss Apotome microscope. Scale bar represents 1mm for whole colony images and 5µm for single cell images. (F) Cells from colonies were isolated and single cells were imaged using Zeiss Apotome microscope and percentage of pseudohyphal cells were calculated using ImageJ. More than 500 cells were counted for each strains. Statistical analysis was done using one-way ANOVA test, **(P<0.01) and *(P<0.05). Error bars represent SEM. (G) Growth curve was performed to monitor overall growth of wild-type strains on SLAD, SLAD+2DG and SLAD+NaCi. Overnight grown culture of wild-type strains (∑1278b or CEN.PK) were diluted to OD600=0.01 in fresh SLAD medium with and without 2DG (0.05%) or NaCi (0.5%) and allowed to grow at 30 °C for 24 hours. OD600 was recorded at 3-h intervals (H) Growth curve was performed to monitor overall growth of wild-type and knockout strains on SLAD. Overnight grown culture of deletion strains (ΔΔpfk1 or ΔΔadh1) along with wild-type strains (∑1278b or CEN.PK) were diluted to OD600=0.01 in fresh SLAD and allowed to grow at 30 °C for 24 hours. OD600 was recorded at 3-h intervals. (I) Schematic overview of the role of glucose and glycolytic intermediates in the activation of cAMP-PKA pathway during pseudohyphal differentiation. (J) Wild-type ∑1278b was spotted on SLAD containing sub-inhibitory concentration of 2DG or NaCi with and without cAMP (1mM), incubated for 10 days at 30 °C. After 10 days, whole colony imaging was done using Olympus MVX10 stereo microscope. Scale bar represents 1mm for whole colony images. (K) Pertinent knockout strains which lack genes that encode for enzymes involved in glycolysis such as pfk1 and adh1 were spotted on SLAD with and without cAMP (1mM), incubated for 10 days at 30 °C. ΔΔgpa2 strain was used as a control. After 10 days, whole colony imaging was done using Olympus MVX10 stereo microscope. Scale bar represents 1mm for whole colony images. This figure was created using BioRender.com.
Another diploid S. cerevisiae strain that is known to undergo pseudohyphal differentiation under nitrogen-limiting conditions is the prototrophic CEN.PK strain (Laxman and Tu, 2011). In order to determine if our observations in ∑1278b can be recapitulated in CEN.PK, S. cerevisiae wild-type CEN.PK was spotted on SLAD with and without sub-inhibitory concentrations of 2-Deoxy-D-Glucose (2DG) and sodium citrate (NaCi). Our data clearly demonstrates that 2DG and NaCi strongly inhibit pseudohyphal differentiation (Fig. 1B and C) without compromising the overall growth of S. cerevisiae wild-type CEN.PK as well (Fig. 1G).
Given that, 2DG and NaCi exhibited a strong inhibitory effect on pseudohyphal differentiation under nitrogen-limiting conditions in a strain-independent manner, we wanted to corroborate these observations by using pertinent genetic knockout strains that are compromised in performing glycolysis, in the presence of glucose (Heinisch, 1986; Smith et al., 2004). In order to do this, we generated knockout strains which lack the genes that encode for enzymes that are involved in the glycolysis pathway including pfk1 (phosphofructokinase-1) and adh1 (alcohol dehydrogenase-1) (Bennetzen and Hall, 1982; Foy and Bhattacharjee, 1978) (Fig. 1D). Heinisch observed that a null mutant of pfk1, a key glycolytic enzyme, is viable and can grow on glucose as a sole carbon source under conditions where cell performs aerobic glycolysis. However, this same mutant exhibits a significant growth defect when cultured anaerobically (Heinisch, 1986; Lobo and Maitra, 1983). Similarly, Dickinson et al., showed that a null mutant of adh1 is viable. This is likely due to the presence of multiple alcohol dehydrogenase isozymes present in S. cerevisiae that likely compensate for the absence of adh1. (Dickinson et al., 2003). The deletions of pfk1 and adh1 genes were confirmed through whole genome sequencing (Suppl.Fig. 1A and B). Our result demonstrates that the deletion of pfk1 and adh1 in the ∑1278b background (ΔΔpfk1 refers to the pfk1 knockout strain and ΔΔadh1 refers to the adh1 knockout strain) resulted in a significant reduction in pseudohyphal differentiation, compared to the corresponding wild-type strain, under nitrogen-limiting conditions (Fig. 1E). We then isolated cells from these colonies and quantified the percentage of pseudohyphal cells in these colonies. Deletion of pfk1 and adh1 resulted in a significant reduction in pseudohyphal cells (Fig. 1F) corroborating our data obtained using the glycolytic inhibitors, 2DG and NaCi. Similarly, deletion of pfk1 and adh1 in the CEN.PK background resulted in a significant reduction in pseudohyphal differentiation (Fig. 1E and F). In order to rule out whether the reduction in pseudohyphal differentiation in these mutants, is due to a general growth defect, we performed growth curve analysis wherein overnight grown cultures of the deletion strains (ΔΔpfk1 and ΔΔadh1) along with wild-type strains (∑1278b or CEN.PK) were diluted to OD600=0.01 in fresh SLAD and allowed to grow at 30 °C for 24 hours. OD600 was recorded at 3-h intervals (Fig. 1H). Our data suggests that the reduction in pseudohyphal differentiation is not due to general growth defects as the mutants were able to grow similar to the wild-type strains in SLAD (Fig. 1H). Taken together, our results clearly demonstrate that the ability of S. cerevisiae to efficiently metabolize glucose via glycolysis is critical for pseudohyphal differentiation under nitrogen-limiting conditions.
It is well-established that in S. cerevisiae, glucose acts as a key signalling molecule and is sensed by the G-protein-coupled receptor Gpr1 that activates the cAMP-PKA pathway, which in turn is crucial for pseudohyphal differentiation under nitrogen-limiting conditions (Lorenz and Heitman, 1997; Lorenz et al., 2000). Additionally, a key glycolytic intermediate, fructose-1,6-bisphosphate (FBP), is also known to activate the cAMP-PKA pathway via activation of Ras proteins, further linking glucose metabolism to this crucial signalling cascade (Peeters et al., 2017) (Fig. 1I). To investigate whether the pseudohyphal differentiation defects that we observed due to glycolysis perturbation were solely due to the inability of glucose to activate the cAMP-PKA pathway, we performed cAMP add-back assays wherein S. cerevisiae wild-type ∑1278b was spotted on SLAD containing sub-inhibitory concentrations of 2DG or NaCi in the presence and absence of 1mM cAMP (Lorenz and Heitman, 1997). Interestingly, the exogenous addition of cAMP failed to rescue pseudohyphal differentiation defect caused by the perturbation of glycolysis through 2DG or NaCi under nitrogen-limiting condition (Fig. 1J). We next asked if the exogenous addition of cAMP rescues the pseudohyphal differentiation defect exhibited by glycolysis defective strains, ΔΔpfk1 and ΔΔadh1. In order to do this, we performed similar cAMP add-back assays wherein S. cerevisiae wild-type ∑1278b along with ΔΔpfk1 and ΔΔadh1 strains were spotted on SLAD in the presence and absence of 1mM cAMP. The ΔΔgpa2 strain was used as a control because it is well-established that the deletion of gpa2 (G protein that activates cAMP production in response to glucose (Rolland et al., 2000)) impairs pseudohyphal differentiation and this defect can be fully rescued by the exogenous addition of cAMP under nitrogen-limiting conditions (Harashima and Heitman, 2002; Lorenz and Heitman, 1997). Corroborating our previous data, the exogenous addition of cAMP failed to rescue pseudohyphal differentiation defect caused by the perturbation of glycolysis via the deletion of pfk1 and adh1 but fully rescued the pseudohyphal differentiation defect exhibited by the ΔΔgpa2 strain (Fig. 1K). Taken together, our results clearly demonstrate that the ability of S. cerevisiae to efficiently metabolize glucose via glycolysis is critical for pseudohyphal differentiation in a cAMP-PKA pathway independent manner under nitrogen-limiting conditions. This implies that glycolysis may be parallelly regulating other cellular processes essential for this morphological transition under nitrogen-limiting conditions.
Comparative Transcriptomics Identifies Glycolysis-dependent Regulation of Sulfur Metabolism During Pseudohyphal Differentiation in S. cerevisiae
Our previous data showed that the perturbation of glycolysis using specific inhibitors including 2DG/NaCi or deletion of genes that encode for glycolytic enzymes including pfk1/ adh1 resulted in a significant reduction in pseudohyphal differentiation in a cAMP-PKA independent manner, under nitrogen-limiting conditions (Fig. 1). Given that, glycolysis strongly influences the global transcriptional landscape of S. cerevisiae (Newcomb et al., 2003; Wu et al., 2019), we undertook a comprehensive transcriptomic approach to identify the changes in the transcriptional landscape of S. cerevisiae colonies, in conditions that perturb glycolysis, resulting in the inhibition of pseudohyphal differentiation. We performed comparative RNA-Seq analysis using RNA isolated from wild-type cells spotted on SLAD in the presence and absence of the glycolysis inhibitor, 2DG (2DG addition was used to perturb glycolysis in our transcriptomics experiments, as it exerted the strongest inhibitory effect on pseudohyphal differentiation as shown in Fig. 1B and C). The experimental schematic is depicted in Fig. 2A. Briefly, comparative RNA-Seq analysis was done using wild-type colonies grown on SLAD in the presence and absence of 2DG isolated on day 5 (D-5) and day 10 (D-10), respectively (Fig. 2A). We have used D-5 as an early time point for our experiments, as we have consistently observed the emergence of pseudohyphae in colonies spotted on SLAD at this time-point and D-10 as a late time point wherein colonies exhibit robust pseudohyphal differentiation (Suppl.Fig. 2A). A volcano plot of differential gene expression in 2DG treated cells compared to untreated cells in both D-5 and D-10 is shown in Fig. 2B wherein some of the significantly upregulated and downregulated genes are highlighted with green colour. Previous studies in the field have shown that, under nitrogen-limiting/glucose-replete conditions, glycolytic intermediates provide various precursors needed for the biosynthesis of several amino acids in order to maintain amino acid homeostasis (Dikicioglu et al., 2011; Martíez-force and Benítez, 1992) and perturbation of glycolysis under these conditions results in the increased transcription of various genes involved in amino acid biosynthesis, as a feedback response. As expected, the expression of several genes involved in amino acid biosynthesis and transport were significantly upregulated in 2DG treated cells compared to untreated cells, in both D-5 and D-10 respectively (Suppl.Fig. 2B) corroborating the observations from the previously mentioned studies (Dikicioglu et al., 2011; Wu et al., 2004). Interestingly and rather unexpectedly, the expression of multiple genes specifically involved in the biosynthesis and transport of sulfur-containing amino acids, cysteine and methionine were significantly downregulated in 2DG treated cells compared to untreated cells in both D-5 and D-10 respectively (Fig.2C). Our RNA-Seq analysis clearly demonstrates that multiple genes involved in the de novo biosynthesis of sulfur-containing amino acids, are significantly downregulated in the presence of 2DG, under nitrogen-limiting conditions. In order to determine whether these observed trends are reflected at the protein level, we generated strains wherein various proteins involved in the de novo biosynthesis of sulfur-containing amino acids including Met4 (Leucine-zipper transcriptional activator), Met32 (Zinc-finger DNA-binding transcription factor), Met16 (3’-phosphoadenylsulfate reductase), Met10 (Subunit alpha of assimilatory sulfite reductase), Cys4 (Cystathionine beta-synthase) and Cys3 (Cystathionine gamma-lyase) were epitope-tagged (Fig. 2D). These strains were spotted on SLAD in the presence and absence of 2DG following which, cells from colonies were isolated on D-5 and levels of these proteins were assessed by using Western blotting (Fig. 2E). Our results indicate that the levels of Met4, Met32, Met16, Met10, Cys4 and Cys3 were significantly reduced in the 2DG treated cells compared to untreated cells, corroborating our transcriptome data (Fig. 2E and Suppl.Fig. 3A). This clearly demonstrates that perturbation of glycolysis negatively influences the expression of sulfur metabolism proteins, under nitrogen-limiting conditions. We next asked, if the glycolysis-dependent regulation of sulfur metabolism under nitrogen-limiting conditions (SLAD) is restricted to conditions wherein cells physically attach to solid surfaces (colony growth) or whether it is conserved even in conditions where cells are free-floating (liquid-culture growth). In order to do this, pertinent epitope-tagged strains (wherein Met32, Met16, Met10 and Cys3 were HA-tagged) were grown in liquid SLAD in the presence and absence of 2DG for 24 hours after which the expression of the aforesaid proteins were assessed using Western blotting (Suppl.Fig. 3B and C). Importantly, our results indicate that the levels of Met32, Met16, Met10 and Cys3 were significantly reduced in the 2DG treated cells compared to untreated cells (Suppl.Fig. 3B and C). Taken together, our data clearly demonstrates that the perturbation of glycolysis under nitrogen-limiting conditions negatively influences sulfur metabolism.

Comparative Transcriptomics Identifies the Role of Glycolysis in Regulating Sulfur Metabolism During Pseudohyphal Differentiation in S. cerevisiae.
(A) Schematic overview of steps involved in RNA isolation and RNA-sequencing. (B) Volcano plots represent differentially expressed genes in 2DG treated cells compared to untreated cells isolated from D-5 and D-10 colonies, respectively. Genes which are upregulated (Log2 fold change ≥ 1) and downregulated (Log2 fold change ≤ −1) are highlighted in red. Some of the significantly upregulated and downregulated genes involved in the biosynthesis of various amino acids are highlighted in green. (C) Heatmaps represent the differentially expressed genes involved in de novo biosynthesis and transport of sulfur-containing amino acids, in 2DG treated cells compared to untreated cells isolated from D-5 and D-10 colonies, respectively (n=3). Scale bar represents Z-Score. (D) Schematic overview of de novo biosynthesis of sulfur-containing amino acids. (E) Schematic overview of steps involved in the Western blotting experiments. Epitope-tagged strains of various proteins involved in the de novo biosynthesis of sulfur-containing amino acids including Met4, Met32, Met16, Met10, Cys4 and Cys3 were spotted on SLAD in the presence and absence of sub-inhibitory concentration of 2DG following which, cells from colonies were isolated on D-5 and levels of these proteins were assessed using Western blotting (n=3). β-actin was used as loading control. This figure was created using BioRender.com.

Glycolysis-mediated Regulation of Sulfur Metabolism is Critical for Pseudohyp-hal Differentiation in S. cerevisiae.
(A) Wild-type ∑1278b was spotted on SLAD containing sub-inhibitory concentration of 2DG with and without sulfur-containing compounds including cysteine (500µM) or methionine (20µM), incubated for 10 days at 30 °C. After 10 days, whole colony imaging was done using Olympus MVX10 stereo microscope. Scale bar represents 1mm for whole colony images. (B) Cells from colonies were isolated and single cells were imaged using Zeiss Apotome microscope and percentage of pseudohyphal cells were calculated using ImageJ. More than 500 cells were counted for each condition. Statistical analysis was done using unpaired t-test, ***(P<0.001) and ns (non-significant). Error bars represent SEM. (C) Pertinent knockout strains which lack genes encoding for enzymes involved in glycolysis such as pfk1 and adh1 along with wild-type were spotted on SLAD with and without sulfur-containing compounds including cysteine (200µM) or methionine (20µM), incubated for 10 days at 30 °C. After 10 days, whole colony imaging was done using Olympus MVX10 stereo microscope. Scale bar represents 1mm for whole colony images. (D) Cells from colonies were isolated and single cells were imaged using Zeiss Apotome microscope and percentage of pseudohyphal cells were calculated using ImageJ. More than 500 cells were counted for each condition. Statistical analysis was done using one-way ANOVA test, ****(P<0.0001), ***(P<0.001), **(P<0.01) and *(P<0.05). Error bars represent SEM. (E) Pertinent knockout strains which lack genes encoding for transcription factor involved in de novo biosynthesis of sulfur-containing amino acids including met32 along with wild-type ∑1278b were spotted on SLAD and imaged after 10 days using Olympus MVX10 stereo microscope. Scale bar represents 1mm for whole colony images. Cells from colonies were isolated and single cells were imaged using Zeiss Apotome microscope and percentage of pseudohyphal cells were calculated using ImageJ. More than 500 cells were counted for each condition. Statistical analysis was done using unpaired t-test, ***(P<0.001). Error bars represent SEM. (F) Pertinent knockout strain ΔΔmet32 was spotted on SLAD with and without sulfur-containing compounds including cysteine (200µM) or methionine (20µM), incubated for 10 days at 30 °C. Wild-type spotted on SLAD was used as control. After 10 days, imaging was done using Olympus MVX10 stereo microscope. Scale bar represents 1mm for whole colony images. Cells from colonies were isolated and single cells were imaged using Zeiss Apotome microscope and percentage of pseudohyphal cells were calculated using ImageJ. More than 500 cells were counted for each condition. Statistical analysis was done using one-way ANOVA test, ***(P<0.001) and **(P<0.01). Error bars represent SEM. (G) Schematic overview of the proposed role of SCFMet30 in glycolysis-dependent regulation of sulfur metabolism during pseudohyphal differentiation under nitrogen-limiting conditions in S. cerevisiae. (H) Epitope-tagged strains of various proteins involved in the de novo biosynthesis of sulfur-containing amino acids including Met4, Met32, Met16 and Cys3 in Δmet30 background were spotted on SLAD in the presence and absence of sub-inhibitory concentration of 2DG following which, cells from colonies were isolated on D-5 and levels of these proteins were assessed using Western blotting (n=3). β-actin was used as loading control. (I) Raw images of Western blots were analysed using ImageJ to normalize targeted protein expression with the expression of housekeeping protein-β-actin, in order to generate densitometric graphs. Statistical analysis was done using unpaired t-test, ns (non-significant). Error bars represent SEM. (J) Epitope-tagged strain of Met30 was spotted on SLAD in the presence and absence of sub-inhibitory concentration of 2DG following which, cells from colonies were isolated on D-5 and levels of these proteins were assessed using Western blotting (n=3). β-actin was used as loading control. Raw images of Western blots were analysed using ImageJ to normalize targeted protein expression with the expression of housekeeping protein-β-actin, in order to generate densitometric graphs. Statistical analysis was done using unpaired t-test, ns (non-significant). Error bars represent SEM. This figure was created using BioRender.com.
Exogenous Supplementation of Sulfur Sources Rescues Pseudohyphal Differentiation Defects Caused by the Perturbation of Glycolysis in S. cerevisiae
Our comparative transcriptomics data clearly demonstrates that, under nitrogen-limiting conditions, glycolysis plays a critical role in regulating the expression of genes involved in the de novo biosynthesis of sulfur-containing amino acids (Fig. 2). Based on our previous data, we hypothesized that under nitrogen-limiting conditions, active glycolysis enables de novo biosynthesis of sulfur-containing amino acids which in turn, is critical for pseudohyphal differentiation and perturbation of active glycolysis using inhibitors like 2DG or disruption of genes involved in glycolysis including pfk1 or adh1 negatively affects glycolysis-mediated regulation of sulfur metabolism leading to attenuation of pseudohyphal differentiation. If the proposed hypothesis holds true, the exogenous supplementation of sulfur-containing compounds should mitigate the fungal morphogenetic defects resulting from the perturbation of glycolysis. In order to test this, we performed sulfur add-back assays wherein S. cerevisiae wild-type ∑1278b was spotted on SLAD containing sub-inhibitory concentration of 2DG in the presence and absence of either sulfur-containing amino acids (cysteine or methionine). Remarkably, the exogenous addition of cysteine resulted in a significant rescue of the pseudohyphal differentiation defect observed in nitrogen-limiting medium containing 2DG (Fig. 3A). Corroborating this observation, the percentage of pseudohyphal cells in colonies isolated from SLAD containing 2DG and cysteine were significantly higher than the percentage of pseudohyphal cells in colonies isolated from SLAD containing 2DG alone (Fig. 3B), Interestingly, the addition of methionine failed to rescue the pseudohyphal differentiation defects caused by the perturbation of glycolysis (Fig. 3A and B).
We next asked if the exogenous addition of sulfur sources rescues the pseudohyphal differentiation defect exhibited by glycolysis defective strains, ΔΔpfk1 and ΔΔadh1. In order to do this, we performed similar sulfur add-back assays wherein S. cerevisiae wild-type ∑1278b along with ΔΔpfk1 and ΔΔadh1 strains were spotted on SLAD in the presence and absence of either sulfur-containing amino acids (cysteine or methionine). The exogenous addition of both cysteine or methionine was able to completely rescue the pseudohyphal differentiation defect exhibited by the ΔΔpfk1 and ΔΔadh1 strains (Fig. 3C). Corroborating this observation, the percentage of pseudohyphal cells in ΔΔpfk1 and ΔΔadh1 colonies isolated from SLAD containing cysteine or methionine were significantly higher than the percentage of pseudohyphal cells in ΔΔpfk1 and ΔΔadh1 colonies isolated from SLAD alone (Fig. 3D). Taken together, our results clearly demonstrate that the glycolysis-mediated regulation of sulfur metabolism is critical for pseudohyphal differentiation under nitrogen-limiting conditions.
Sulfur Metabolism is Critical for Pseudohyphal Differentiation Under Nitrogen-limiting Conditions in S. cerevisiae
Our previous observations strongly suggest a direct role for sulfur metabolism in regulating fungal morphogenesis, under nitrogen-limiting conditions. In order to determine whether glycolysis-dependent sulfur metabolism is critical for pseudohyphal differentiation, under nitrogen-limiting conditions, we generated a knockout strain that lacks met32 (gene that encodes for a zinc-finger DNA-binding transcription factor involved in the regulation of de novo biosynthesis of sulfur-containing amino acids). Unlike deletions of most other genes involved in the de novo biosynthesis of sulfur-containing amino acids, met32 deletion strain does not exhibit auxotrophy for methionine or cysteine (Blaiseau et al., 1997). This is likely due to the functional redundancy provided by its paralog, met31 (Blaiseau et al., 1997). We then assessed the ability of the met32 deletion strain to undergo pseudohyphal differentiation under nitrogen-limiting conditions. Deletion of met32 resulted in a significant reduction in pseudohyphal differentiation. Corroborating this observation, the percentage of pseudohyphal cells in ΔΔmet32 colonies were significantly lower in SLAD, compared to S. cerevisiae wild-type ∑1278b colonies (Fig. 3E). This suggests that perturbations to sulfur metabolism negatively affects pseudohyphal differentiation. In order to rule out whether the reduction in pseudohyphal differentiation in this mutant, is due to a general growth defect, we performed growth curve analysis wherein overnight grown cultures of deletion strain (ΔΔmet32) along with wild-type strain (∑1278b) were diluted to OD600=0.01 in fresh SLAD and allowed to grow at 30 °C for 24 hours. OD600 was recorded at 3-h intervals (Suppl.Fig. 4A). Our data suggests that the reduction in pseudohyphal differentiation is not due to general growth defect as the mutant was able to grow similar to the wild-type strain in SLAD. We next asked if the exogenous addition of sulfur sources rescues the pseudohyphal differentiation defect exhibited by ΔΔmet32. In order to do this, we performed a simple sulfur add-back assay, wherein ΔΔmet32 strain was spotted on SLAD containing either cysteine or methionine and allowed to undergo pseudohyphal differentiation. Our results show that the addition of cysteine or methionine completely rescued the pseudohyphal differentiation defect exhibited by ΔΔmet32 strain. The percentage of pseudohyphal cells were also significantly higher in ΔΔmet32 colonies growing on SLAD containing cysteine or methionine compared to just SLAD (Fig. 3F). Taken together, our data clearly demonstrates that the efficient de novo biosynthesis of sulfur-containing amino acids including cysteine and methionine is critical for pseudohyphal differentiation, under nitrogen-limiting conditions.

Glycolysis-dependent Sulfur Metabolism is Critical for Hyphal Differentiation in C. albicans.
(A) Schematic overview of glycolysis and its inhibition by glycolysis inhibitor (2-Deoxy-D-Glucose (2DG) (B) Wild-type SC5314 was spotted on SLAD with and without sub-inhibitory concentration of glycolysis inhibitor 2DG, incubated for 7 days at 37 °C. Cells from colonies were isolated and single cells were imaged using Zeiss Apotome microscope and percentage of hyphal cells were calculated using ImageJ. More than 500 cells were counted for each condition. Scale bar represents 10µm for single cell images. Statistical analysis was done using unpaired t-test, **(P<0.01). Error bars represent SEM. (C) Schematic overview of de novo biosynthesis of sulfur-containing amino acids. (D) Wild-type SC5314 was spotted on SLAD, in the presence and absence of sub-inhibitory concentration of 2DG and cells from colonies were isolated after ∼4 days, for RNA isolation. We then performed comparative RT-qPCR to check for the relative expression of genes involved in the de novo biosynthesis of sulfur-containing amino acids pathway including met32, met3, met5 (ecm17), met10 and met17 (met15). Statistical analysis was done using one-way ANOVA test, ****(P<0.0001) and ***(P<0.001). Error bars represent SEM. (E) Wild-type SC5314 was spotted on SLAD containing sub-inhibitory concentration of 2DG with and without cysteine (100µM) and, incubated for 7 days at 37 °C. Cells from colonies were isolated and single cells were imaged using Zeiss Apotome microscope and percentage of hyphal cells were calculated using ImageJ. Scale bar represents 10µm for single cell images. More than 500 cells were counted for each condition. Statistical analysis was done using unpaired t-test, **(P<0.01). Error bars represent SEM. (F) Wild-type SC5314 was spotted on SLAD containing sub-inhibitory concentration of 2DG with and without methionine (20µM) and, incubated for 7 days at 37 °C. Cells from colonies were isolated and single cells were imaged using Zeiss Apotome microscope and percentage of hyphal cells were calculated using ImageJ. More than 500 cells were counted for each condition. Scale bar represents 10µm for single cell images. Statistical analysis was done using unpaired t-test, **(P<0.01). Error bars represent SEM. (G) Wild-type SC5314 and pertinent knockout strain which lacks the genes encoding for a glycolytic enzyme, pfk1 (ΔΔpfk1) were spotted on SLAD, incubated for 7 days at 37 °C. Cells from colonies were isolated and single cells were imaged using Zeiss Apotome microscope and percentage of hyphal cells were calculated using ImageJ. Scale bar represents 10µm for single cell images. More than 500 cells were counted for each strain. Statistical analysis was done using unpaired t-test, **(P<0.01). Error bars represent SEM. (H) Growth curve was performed to monitor overall growth of wild-type strain on SLAD and SLAD+2DG and ΔΔpfk1 on SLAD. Overnight grown culture of wild-type strain (SC5314) were diluted to OD600=0.01 in fresh SLAD medium with and without 2DG (0.2%) and allowed to grow at 30 °C for 24 hours. OD600 was recorded at 3-h intervals. For ΔΔpfk1, overnight grown culture of ΔΔpfk1 strains along with wild-type strain (SC5314) were diluted to OD600=0.01 in fresh SLAD and allowed to grow at 30 °C for 24 hours. OD600 was recorded at 3-h intervals. (I) ΔΔpfk1 along with Wild-type SC5314 was spotted on SLAD and cells from colonies were isolated after ∼4 days, for RNA isolation. We then performed comparative RT-qPCR to check for the relative expression of genes involved in the de novo biosynthesis of sulfur-containing amino acids pathway including met32, met3, met5 (ecm17), met10 and met17 (met15). Statistical analysis was done using one-way ANOVA test, ****(P<0.0001), ***(P<0.001), **(P<0.01) and *(P<0.05). Error bars represent SEM. (J) ΔΔpfk1 was spotted on SLAD in the presence and absence of cysteine (100µM), incubated for 7 days at 37 °C. Cells from colonies were isolated and single cells were imaged using Zeiss Apotome microscope and percentage of hyphal cells were calculated using ImageJ. More than 500 cells were counted for each condition. Statistical analysis was done using unpaired t-test, **(P<0.01). Error bars represent SEM. (K) ΔΔpfk1 was spotted on SLAD in the presence and absence of methionine (10µM), incubated for 7 days at 37 °C. Cells from colonies were isolated and single cells were imaged using Zeiss Apotome microscope and percentage of hyphal cells were calculated using ImageJ. More than 500 cells were counted for each condition. Statistical analysis was done using unpaired t-test, **(P<0.01). Error bars represent SEM. This figure was created using BioRender.com.
Met30 is Involved in the Glycolysis-dependent Regulation of Sulfur Metabolism During Pseudohyphal Differentiation in S. cerevisiae
Our previous data clearly demonstrates that, under nitrogen-limiting conditions, the levels of Met4, a key activator of genes involved in the de novo biosynthesis of sulfur-containing amino acids, was significantly reduced in 2DG treated cells compared to untreated cells (Fig. 2E). Met4 activity is known to be negatively regulated by SCFMet30 ubiquitin ligase complex in S. cerevisiae (Rouillon et al., 2000). Under sulfur-replete conditions, Met30 promotes Met4 ubiquitination and degradation, thereby repressing the Met gene network. Conversely, sulfur limitation reduces Met30 activity, allowing Met4 to become active and induce the expression of genes required for sulfur amino acid biosynthesis (Smothers et al., 2000). It is important to note that deletion of met30 is lethal in haploid strains of S. cerevisiae (Kaiser et al., 1998). While point mutations or temperature sensitive met30 mutants are commonly used in haploid strains to circumvent this lethality (Kaiser et al., 1998; Su et al., 2008), our study in diploid S. cerevisiae involved generating a heterozygous viable met30 (Δmet30) deletion strain. Based on our previous data, we hypothesized that under nitrogen-limiting conditions, active glycolysis enables de novo biosynthesis of sulfur-containing amino acids by reducing the activity of Met30. Conversely, we hypothesized that perturbing active glycolysis with inhibitors like 2DG increases Met30 activity, which in turn inactivates Met4, thereby negatively affecting Met gene transcription, ultimately resulting in attenuated pseudohyphal differentiation (Fig. 3G). In order to investigate this hypothesis, we generated strains wherein one copy of met30 was deleted and various proteins involved in the de novo biosynthesis of sulfur-containing amino acids including Met4 (Leucine-zipper transcriptional activator), Met32 (Zinc-finger DNA-binding transcription factor), Met16 (3’-phosphoadenylsulfate reductase) and Cys3 (Cystathionine gamma-lyase) were epitope-tagged. These strains were spotted on SLAD in the presence and absence of 2DG following which, cells from colonies were isolated on D-5 and levels of these proteins were assessed by using Western blotting. Our results indicate that the levels of Met4, Met32, Met16 and Cys3 were similar in both 2DG treated and untreated cells, in the Δmet30 deletion strain (Fig. 3H and I). This suggests that glycolysis influences Met30 activity which in turn regulates Met4 activity thereby affecting the expression of genes involved in the de novo biosynthesis of sulfur-containing amino acids, under nitrogen-limiting conditions.
Based on our previous observation that 2DG treatment significantly reduces Met4 levels in nitrogen-limiting conditions, (an effect that was notably absent in met30 deletion strains), we hypothesized that the 2DG-induced downregulation of Met4 is mediated by an increase in Met30 levels. In order to check Met30 levels in the presence of 2DG, we generated a strain wherein Met30 was epitope-tagged and this strain was spotted on SLAD, in the presence and absence of 2DG following which, cells from colonies were isolated on D-5 and levels of Met30 were assessed by using Western blotting. Our results indicate that the levels of Met30 were similar in both 2DG treated and untreated cells (Fig. 3J), suggesting that glycolysis influences Met30 activity post-translationally rather than through changes in its protein abundance. Collectively, our findings indicate that the perturbation of glycolysis under nitrogen-limiting conditions negatively impacts sulfur metabolism through a mechanism dependent on Met30 activity.
Glycolysis-dependent Sulfur Metabolism is Critical for Hyphal Differentiation in C. albicans
C. albicans is an opportunistic fungal pathogen that exhibits a remarkable degree of fungal morphogenesis in response to various environmental cues (Huang, 2012). Yeast to hyphal differentiation in C. albicans is well studied. However, we do not have a complete understanding of fundamental metabolic drivers that orchestrate this phenomenon. Given that central carbon metabolic pathways including glycolysis are well conserved across a plethora of fungal species including C. albicans (Askew et al., 2009) and nitrogen limitation is one of the key triggers for yeast to hyphal differentiation in C. albicans (Csank et al., 1998), we wanted to determine if glycolysis-mediated sulfur metabolism plays an important role in this morphogenetic switching in C. albicans as well. First, we wanted to understand whether the ability of C. albicans cells to metabolize glucose, under nitrogen-limiting conditions is critical for fungal morphogenesis (Fig. 4A). In order to do this, C. albicans wild-type SC5314 was spotted on SLAD, in the presence and absence of sub-inhibitory concentration of glycolysis inhibitor, 2DG, allowed to grow for 7 days, and cells were isolated and imaged using bright-field microscopy. Our data clearly demonstrates that sub-inhibitory concentration of 2DG significantly reduced filamentation in C. albicans (Fig. 4B) without compromising the overall growth of C. albicans wild-type SC5314 (Fig. 4H). We then isolated cells from these colonies and quantified the percentage of hyphal cells (cells with a length/width ratio of 4.5 or more (Su et al., 2018)) in these colonies. Corroborating our microscopy data, the percentage of hyphal cells in colonies spotted on SLAD containing 2DG was significantly lower compared to the percentage of hyphal cells in colonies spotted on SLAD without 2DG (Fig. 4B). Taken together, our results show that the ability of C. albicans to metabolize glucose is critical for fungal morphogenesis.
Given that the perturbation of glycolysis under nitrogen-limiting conditions, using 2DG exhibited a strong inhibitory effect on hyphal differentiation in C. albicans (Fig. 4B), similar to what we observed in S. cerevisiae, we wanted to determine if the expression of genes involved in the de novo biosynthesis of sulfur-containing amino acids are affected when glycolysis is perturbed in C. albicans, under nitrogen-limiting conditions (similar to what we observed in S. cerevisiae) (Chebaro et al., 2017). In order to determine this, we spotted C. albicans wild-type SC5314 on SLAD, in the presence and absence of 2DG and cells from colonies were isolated after ∼4 days for RNA isolation. We then performed comparative RT-qPCR to check for the relative expression of genes involved in de novo biosynthesis of sulfur-containing amino acids including met32 (DNA-binding transcription factor), met3 (ATP sulfurylase), met5 (ecm17) (Sulfite reductase beta subunit), met10 (Sulfite reductase), and met17 (met15) (O-acetyl homoserine-O-acetyl serine sulfhydrylase) (Chebaro et al., 2017; Li et al., 2013; Shrivastava et al., 2021), in cells isolated from colonies grown on SLAD containing 2DG compared to cells isolated from colonies grown on SLAD without 2DG (Fig. 4C). Our data clearly shows that all the aforesaid genes involved in de novo biosynthesis of sulfur-containing amino acids pathway including met32, met3, met5 (ecm17), met10 and met17 (met15) were significantly downregulated in the presence of 2DG (Fig. 4D) suggesting that perturbation of glycolysis negatively impacts sulfur metabolism in C. albicans, under nitrogen-limiting conditions. To determine if the significant reduction in hyphal differentiation that we observed in the presence of 2DG can be attributed to the reduced expression of genes involved in the de novo biosynthesis of sulfur-containing amino acids, we performed sulfur add-back assays similar to what was done in S. cerevisiae (Fig. 3A), wherein C. albicans wild-type SC5314 was spotted on SLAD containing 2DG in the presence and absence of cysteine and methionine, allowed to grow for 7 days, and cells were isolated and imaged using bright-field microscopy. The addition of cysteine and methionine significantly rescued hyphal differentiation defects in the presence of 2DG (Fig. 4E and F). We then isolated cells from these colonies and quantified the percentage of hyphal cells. Corroborating our microscopy data, the percentage of hyphal cells in colonies spotted on SLAD containing 2DG with cysteine or methionine were significantly higher compared to the percentage of hyphal cells in colonies spotted on SLAD with 2DG alone (Fig. 4E and F).
Our previous results clearly demonstrate that glycolysis and glycolysis-dependent sulfur metabolism is critical for fungal morphogenesis of C. albicans, under nitrogen-limiting conditions. In order to corroborate these observations by using pertinent genetic knockout strains that are compromised in performing glycolysis, we generated a double knockout strain lacking the gene that encodes for the glycolytic enzyme, phosphofructokinase (Pfk1) using the SAT-Flipper method (Reuß et al., 2004). The deletion of pfk1 was confirmed through whole genome sequencing (Suppl.Fig. 5A). We then tested the ability of this deletion strain to undergo hyphal differentiation under nitrogen-limiting conditions. Briefly, C. albicans wild-type SC5314 and ΔΔpfk1 were spotted on SLAD, allowed to grow for 7 days, and cells were isolated and imaged using bright-field microscopy. Our result shows that the ΔΔpfk1 strain exhibits a significant reduction in hyphal differentiation compared to C. albicans wild-type SC5314, under nitrogen-limiting conditions (Fig. 4G). We then isolated cells from these colonies and quantified the percentage of hyphal cells in these colonies. Corroborating our microscopy data, the percentage of hyphal cells in C. albicans wild-type SC5314 colonies spotted on SLAD were significantly higher compared to the percentage of hyphal cells in ΔΔpfk1 colonies (Fig. 4G). In order to rule out whether the reduction in hyphal differentiation in this mutant, is due to a general growth defect, we performed growth curve analysis wherein overnight grown cultures of deletion strain (ΔΔpfk1) along with wild-type strain (SC5314) were diluted to OD600=0.01 in fresh SLAD and allowed to grow at 30 °C for 24 hours. OD600 was recorded at 3-h intervals.Our data suggests that the reduction in hyphal differentiation is not due to general growth defects as the mutant was able to grow similar to the wild-type strain in SLAD (Fig. 4H).

Perturbation of Glycolysis Leads to Attenuated Fungal Virulence in C. albicans.
(A) Schematic overview of various proteins involved in hyphal differentiation and virulence of C. albicans. (B) Wild-type SC5314 was spotted on SLAD, in the presence and absence of sub-inhibitory concentration of 2DG or wild-type SC5314 and ΔΔpfk1 were spotted on SLAD and cells from these colonies were isolated after ∼4 days, for RNA isolation. We then performed comparative RT-qPCR to check for the relative expression of genes involved in hyphal differentiation and virulence of C. albicans including als3, ece1, hwp1, hyr1, ihd1, rbt1 and sap6. Statistical analysis was done using one-way ANOVA test, ****(P<0.0001) and ***(P<0.001). Error bars represent SEM. (C) Schematic overview of in vitro microbial survival assay. RAW 264.7 macrophages were incubated with the wild-type and ΔΔpfk1 strain with MOI=10. After 1 hour of incubation, macrophages were lysed and plating was done to enumerate CFU. Percentage of survival was expressed as the number of CFU in the presence of macrophages divided by the number of CFU in the absence of macrophages. Statistical analysis was done using unpaired t-test, ***(P<0.001) and **(P<0.01). Error bars represent SEM. (D) 6-8-week-old C57BL/6 mice were infected intravenously via lateral tail vein injection with 1 × 107 CFU of wild-type SC5314 or 1 × 107 CFU of ΔΔpfk1. To determine survival rate, mice were monitored for 21 days post-infection for clinical signs of illness or mortality. The results are representative of seven mice per injected strain. Survival percentage were statistically evaluated by the Log-rank Mantel-Cox test, ***(P<0.001) (n=7). (E) Schematic overview of kidney isolation to check fungal burden using CFU counting and histology. (F) Fungal burden was measured by CFU counting after kidney isolation from mice injected with either wild-type or ΔΔpfk1. Statistical analysis was done using unpaired t-test, **(P<0.01). Error bars represent SEM. (G) Histopathological analysis of kidney sections using Grocott Methanamine Silver (GMS) staining was done for kidneys isolated from mice injected with either wild-type SC5314 or ΔΔpfk1. Fungal cells appear black in colour against the colored background of the kidney tissue. Bright field imaging was done using Axioplan 2 microscope at 40X magnification. Scale bar represents 50 µm. (H) Schematic overview of N-acetyl cysteine (NAC) administration to mice. (I) 6mg/ml of NAC was dissolved in distilled water and administered orally to 6-8-week-old C57BL/6 mice before 72 hours of C. albicans infection. Mice administered with normal distilled water were used as controls. After 72 hours of treatment, mice were infected intravenously via lateral tail vein injection with 1 × 107 CFU of SC5314 or 1 × 107 CFU of ΔΔpfk1. To determine survival rate, mice were monitored for 21 days post-infection for clinical signs of illness or mortality. The results are representative of six mice per condition. Survival percentage were statistically evaluated by the Log-rank Mantel-Cox test, ***(P<0.001) and **(P<0.01) (n=6). (J) Histopathological analysis of kidney sections using GMS staining was done for kidneys isolated from mice administrated with normal distilled water or NAC containing distilled water and injected with ΔΔpfk1. Fungal cells appear black in colour against the colored background of the kidney tissue. Bright field imaging was done using Axioplan 2 microscope at 40X magnification. Scale bar represents 50 µm. This figure was created using BioRender.com.
In order to determine if the expression of genes involved in the de novo biosynthesis of sulfur-containing amino acids are affected in ΔΔpfk1, under nitrogen-limiting conditions (similar to what we observed in the presence of 2DG (Fig. 4D)), we spotted C. albicans wild-type SC5314 and ΔΔpfk1 on SLAD and cells from colonies were isolated after ∼4 days, for RNA isolation. We then performed comparative RT-qPCR to check for the relative expression of genes involved in de novo biosynthesis of sulfur-containing amino acids including met32 (DNA-binding transcription factor), met3 (ATP sulfurylase), met5 (ecm17) (Sulfite reductase beta subunit), met10 (Sulfite reductase) and met17 (met15) (O-acetyl homoserine-O-acetyl serine sulfhydrylase) (Chebaro et al., 2017; Li et al., 2013; Shrivastava et al., 2021)(Fig. 4C). Our data clearly shows that all the aforesaid genes involved in de novo biosynthesis of sulfur-containing amino acids pathway including met32, met3, met5 (ecm17), met10 and met17 (met15) were significantly downregulated in ΔΔpfk1 compared to the wild-type (Fig. 4I) suggesting that the genetic perturbation of glycolysis negatively impacts sulfur metabolism in C. albicans, under nitrogen-limiting conditions.
Our RT-qPCR results showed that the, deletion of pfk1, negatively influences sulfur metabolism and this in turn might be causal for the reduced hyphal differentiation observed in these conditions. Given this, we wanted to determine if the hyphal differentiation defect exhibited by ΔΔpfk1 can be rescued by the addition of sulfur-containing amino acids. In order to do this, we performed sulfur add-back assays wherein ΔΔpfk1 strain was spotted on SLAD in the presence and absence of cysteine and methionine, allowed to grow for 7 days and imaged using bright field microscopy. Interestingly, our data shows that cysteine and methionine were able to rescue the hyphal differentiation defect exhibited by the ΔΔpfk1 strain (Fig. 4J and K). We then isolated cells from these colonies and quantified the percentage of hyphal cells. Corroborating our microscopy data, the percentage of hyphal cells in ΔΔpfk1 colonies spotted on SLAD containing cysteine and methionine were significantly higher compared to the percentage of hyphal cells in ΔΔpfk1 colonies spotted on SLAD alone (Fig. 4J and K). Overall, our data suggests that glycolysis-mediated regulation of sulfur metabolism is critical for hyphal differentiation of C. albicans, under nitrogen-limiting conditions highlighting the remarkable similarities with respect to the regulation of fungal morphogenesis, between S. cerevisiae and C. albicans.
Perturbation of Glycolysis Leads to Attenuated Fungal Virulence in C. albicans which is Rescued by Sulfur Supplementation
Several studies have established the importance of fungal morphogenesis as a key virulence strategy used by C. albicans to establish successful infections in the host (Gow et al., 2012; Moyes et al., 2016; Wilson et al., 2016). Based on our previous observations that perturbation of glycolysis, either through 2DG treatment or deletion of pfk1 (ΔΔpfk1), significantly attenuates hyphal differentiation, we hypothesized that this observed phenotype could be a consequence of transcriptional downregulation of genes associated with hyphal differentiation and virulence. In order to test this hypothesis, we spotted C. albicans wild-type SC5314 on SLAD, in the presence and absence of 2DG or wild-type along with ΔΔpfk1 on SLAD. Cells from colonies were isolated after ∼4 days for RNA isolation. We then performed comparative RT-qPCR to check for the relative expression of genes involved in hyphal differentiation and virulence including als3 (Cell wall adhesin), ece1 (Candidalysin), hwp1 (Hyphal cell wall protein), hyr1 (GPI-anchored hyphal cell wall protein), ihd1 (GPI-anchored protein), rbt1 (Cell wall protein) and sap6 (Secreted aspartyl protease) (Bailey et al., 1996; Deorukhkar and Roushani, 2017; Liu and Filler, 2011; Martin et al., 2013; Monniot et al., 2013; Moyes et al., 2016; Richardson et al., 2018; Sharkey et al., 1999) (Fig. 5A). Our data clearly shows that all the aforesaid genes involved in hyphal differentiation and virulence including als3, ece1, hwp1, hyr1, ihd1, rbt1 and sap6 were significantly downregulated in both 2DG treated cells and in ΔΔpfk1 suggesting that perturbation of glycolysis negatively affects the expression of genes involved in hyphal differentiation and virulence in C. albicans, under nitrogen-limiting conditions (Fig. 5B).
Given that ΔΔpfk1, displayed a significant attenuation in its ability to undergo hyphal differentiation and it had significantly reduced expression of genes associated with hyphal differentiation and virulence, we wanted to determine if this deletion strain was compromised in its ability to infect the host. In order to do this, we took a two-pronged approach wherein, we first tested the ability of the ΔΔpfk1 strain to survive the harsh intracellular environment of the macrophages which C. albicans encounters during various stages of host infection (Jiménez-López and Lorenz, 2013; May and Casadevall, 2018; Uwamahoro et al., 2014). We performed an in vitro microbial survival assay wherein RAW 264.7 macrophages were incubated either with C. albicans wild-type SC5314 or the ΔΔpfk1 strain for one hour following which macrophages were lysed and fungi were serially diluted and plated to determine the viable population of the C. albicans wild-type SC5314 and the ΔΔpfk1 strain (Fig. 5C). We also calculated the percentage of the wild-type or mutant strain that survived the incubation with the macrophages. Our results demonstrate that the viable population of ΔΔpfk1 strain was significantly lower compared to the wild-type strain, after incubation with the macrophages (Fig. 5C). Similarly, the percentage survival of the ΔΔpfk1 strain was significantly lower than the wild-type strain (Fig. 5C). Overall, this data indicates that the ΔΔpfk1 strain is compromised with respect to macrophage survival.
Given that the ΔΔpfk1 strain was significantly attenuated in the macrophage survival assay, we next wanted to test the ability of this deletion strain to establish systemic infection in a host. In order to do this, we challenged healthy wild-type C57BL/6 mice intravenously with a dose of 1 × 107 CFU (lethal dose (Hirayama et al., 2020)) of C. albicans wild-type SC5314 or ΔΔpfk1 strain via lateral tail vein injection. This is an established murine model of systemic candidiasis in the field and faithfully mimics the infection that occurs in humans (Segal and Frenkel, 2018). We then monitored survival of the infected animals for 21 days. Mice challenged with ΔΔpfk1 strain had significantly increased survival compared to mice challenged with C. albicans wild-type SC5314 (Fig. 5D). C. albicans exhibits strong tropism towards kidneys during murine infections (Lionakis et al., 2011; Pappas et al., 2018) and in order to measure the fungal burden within the host, kidneys from C57BL/6 mice infected with both C. albicans wild-type SC5314 and ΔΔpfk1 were harvested ∼2 days post infection (when the mice injected with wild-type C. albicans SC5314 becomes moribund), homogenized and plated to measure the fungal burden by enumerating the colony forming units (CFU) (Fig. 5E). Our data indicates that kidneys isolated from mice challenged with the ΔΔpfk1 strain had significantly lesser CFU compared to kidneys isolated from mice challenged with the C. albicans wild-type SC5314 (Fig. 5F). We also performed histological analysis on the infected kidneys using GMS staining (Grocott Methenamine Silver staining), to visualize the fungi (stained black) in the tissue. Kidneys isolated from mice challenged with the C. albicans wild-type SC5314 had higher amounts of fungi compared to kidneys isolated from mice challenged with the ΔΔpfk1 strain (Fig. 5G). Overall, our data indicates that the perturbation of glycolysis significantly attenuates C. albicans virulence.
Our previous in vitro results clearly demonstrated that exogenous supplementation of cysteine and methionine significantly rescued the hyphal differentiation defect exhibited by ΔΔpfk1 strain. Based on this, we wanted to check whether sulfur supplementation could also rescue the virulence defects exhibited by ΔΔpfk1, in a murine model of systemic candidiasis. Given that N-acetyl cysteine (NAC) is a commonly used exogenous sulfur source for in vivo studies (Shee et al., 2022), we supplemented distilled water with NAC (6mg/ml concentration) and this was administered to mice orally 72 hours prior to infection. After 72 hours of pre-treatment with NAC, we challenged healthy wild-type C57BL/6 mice intravenously with a dose of 1 × 107 CFU of wild-type SC5314 or ΔΔpfk1 strain via lateral tail vein injection (Fig. 5H). We then monitored survival of the infected animals for 21 days. Remarkably, mice administered with NAC containing distilled water had significantly reduced survival compared to mice administrated with normal distilled water (Fig. 5I). We also performed histological analysis on the infected kidneys using GMS staining, to visualize the fungi in the tissue. Kidneys isolated from mice administered with NAC containing distilled water and infected with ΔΔpfk1 had significantly higher amounts of fungi compared to mice administrated with normal distilled water (Fig. 5J). This clearly demonstrates that exogenous sulfur supplementation to the host via NAC administration, reverses the attenuated virulence exhibited by the ΔΔpfk1 strain.
Discussion
Our findings provide compelling evidence for a conserved metabolic network that intricately links central carbon metabolism (particularly glycolysis), sulfur amino acid biosynthesis, and fungal morphogenesis in both Saccharomyces cerevisiae and Candida albicans. While nitrogen limitation is a well-established and essential trigger for the yeast-to-pseudohyphae transition in S. cerevisiae, the specific influence of carbon sources on this process has not been completely understood. Studies have shown that fermentable carbon sources are critical for inducing pseudohyphal differentiation under nitrogen-limiting conditions (Van De Velde and Thevelein, 2008). Conversely, when these fermentable sugars are replaced with non-fermentable carbon sources, a significant reduction in pseudohyphal differentiation is observed (Strudwick et al., 2010). These findings collectively suggest that the presence of fermentable carbon sources, and by extension their metabolism, is a critical requirement for inducing fungal morphogenesis in S. cerevisiae. Glucose in the extracellular environment can be sensed by S. cerevisiae using multiple proteins. One such protein is the well characterized G-Protein Couple Receptor (GPCR), Gpr1. Binding of glucose to Gpr1 activates a downstream signalling cascade, resulting in the production of cAMP which activates Protein Kinase A (PKA) (Kraakman et al., 1999; Yun et al., 1997). This glucose-dependent cAMP-PKA pathway has been shown to be essential for pseudohyphal differentiation under nitrogen-limiting conditions (Lorenz et al., 2000). Additionally, a key glycolytic intermediate, fructose-1,6-bisphosphate (FBP), is also known to activate the cAMP-PKA pathway via activation of Ras proteins, further linking glucose metabolism to this crucial signalling cascade. However, the role of glucose as a metabolite in the context of pseudohyphal differentiation is not known. Building upon previous data that fermentable carbon sources that are metabolized via glycolysis are able to robustly induce fungal morphogenesis, we asked whether the ability of these cells to undergo active glycolysis is critical for fungal morphogenesis. We took a two-pronged approach which involved pharmacological or genetic perturbation of glycolysis, to address this question. Perturbation of glycolysis using sub-inhibitory concentrations of well-characterized inhibitors (Cramer and Woodward, 1952; Yoshino and Murakami, 1982), 2DG or NaCi significantly attenuated pseudohyphal differentiation in both S. cerevisiae strains (∑1278b and CEN.PK) that are known to undergo pseudohyphal differentiation, without compromising overall growth. This strain-independent effect was further corroborated by genetic evidence wherein deletion of key glycolytic enzyme-encoding genes, pfk1 and adh1, in both the strain backgrounds (∑1278b and CEN.PK) also resulted in a significant reduction in pseudohyphal differentiation, without causing general growth defects. These convergent findings strongly support the notion that active glycolysis is essential for fungal morphogenesis under nitrogen-limiting conditions. To understand whether the pseudohyphal differentiation defects we observed due to glycolysis perturbation were solely due to the inability of glucose to activate glucose-dependent cAMP-PKA pathway, we performed cAMP add-back assays and interestingly, exogenous supplementation of cAMP failed to rescue pseudohyphal differentiation defects caused by the perturbation of glycolysis (through inhibitors (2DG and NaCi) or through genetic knockout strains (ΔΔpfk1 and ΔΔadh1)). Our results clearly demonstrate that the ability of S. cerevisiae to efficiently metabolize glucose via glycolysis is critical for pseudohyphal differentiation under nitrogen-limiting conditions, in a cAMP-PKA independent manner. This implies that glycolysis may be parallelly regulating other cellular mechanisms essential for this morphological transition. It is plausible that the glycolytic flux, or a yet to be identified downstream intermediate, acts as a distinct metabolic signal that is parallelly important for pseudohyphal differentiation along with the cAMP-PKA pathway. This uncovers a previously unrecognized layer of regulatory complexity in the interplay between central carbon metabolism and fungal morphogenesis.
To gain deeper insights underlying the glycolysis-mediated regulation of fungal morphogenesis in a cAMP-PKA independent manner, we performed comparative transcriptomics on S. cerevisiae wild-type cells grown under nitrogen-limiting conditions with and without the glycolysis inhibitor 2DG (which had the strongest inhibitory effect on pseudohyphal differentiation). As anticipated, genes involved in amino acid biosynthesis and transport were upregulated in the 2DG treated cells, consistent with a feedback response that cells exhibits when glycolysis is perturbed under nitrogen-limiting conditions, as various intermediates of glycolysis are known to serve as precursors for the biosynthesis of several amino acids under these conditions (Dikicioglu et al., 2011; Martíez-force and Benítez, 1992). Remarkably, we observed a striking and unexpected downregulation of multiple genes specifically involved in the biosynthesis and transport of sulfur-containing amino acids, cysteine and methionine, in the 2DG treated cells at both early and late time points of pseudohyphal development. This transcriptional downregulation was further validated at the protein level for key proteins involved in the sulfur assimilation pathway (Met4, Met32, Met16, Met10, Cys4 and Cys3), clearly demonstrating the dependence of sulfur metabolism on active glycolysis during fungal morphogenesis. This unexpected link between central carbon metabolism and sulfur metabolism is not restricted to conditions that induce fungal morphogenesis i.e. in surface-attached colonies grown in nitrogen-limiting conditions since even in liquid cultures that have limiting levels of nitrogen, wherein S. cerevisiae cells do not undergo pseudohyphal differentiation (Gancedo, 2001), perturbation of glycolysis with 2DG results in significantly reduced expression of proteins involved in the sulfur assimilation pathway (Met32, Met16, Met10 and Cys3).
Given that multiple genes that encode for proteins involved in de novo biosynthesis of sulfur-containing amino acids were downregulated in the presence of 2DG, which strongly inhibits fungal morphogenesis, we wanted to explore the possibility that perturbation of sulfur metabolism is causal for the attenuation of fungal morphogenesis in the presence of 2DG. We tested the functional relevance of this glycolysis-dependent regulation of sulfur metabolism in fungal morphogenesis directly through exogenous supplementation of sulfur sources. The functional significance of glycolysis-dependent regulation of sulfur metabolism was demonstrated through our sulfur add-back experiments wherein, exogenous supplementation with cysteine specifically rescued the pseudohyphal differentiation defect exhibited by S. cerevisiae in the presence of 2DG. Interestingly, methionine alone was unable to rescue the pseudohyphal differentiation defect, in the presence of 2DG even at various different concentrations (50µM, 100µM, 200µM and 500µM). Furthermore, the pseudohyphal differentiation defects observed in the ΔΔpfk1 and ΔΔadh1 strains were completely rescued by the addition of cysteine or methionine. These findings strongly implicate that de novo biosynthesis of sulfur-containing amino acids is a critical downstream effector of glycolytic activity in promoting pseudohyphal differentiation under nitrogen limitation. To directly assess the role of sulfur metabolism in fungal morphogenesis, we generated and characterized the met32 deletion strain. This particular strain was critical for our investigation because, unlike most other deletions involving genes that encode for proteins involved in the de novo biosynthesis of sulfur-containing amino acids, met32 deletion does not result in auxotrophy for methionine or cysteine. This unique characteristic is attributed to the functional redundancy provided by its paralog, met31 (Blaiseau et al., 1997). The deletion of met32 resulted in substantial defects in pseudohyphal differentiation under nitrogen-limiting conditions, directly implicating the role of sulfur metabolism in regulating fungal morphogenesis. Our sulfur add-back experiments demonstrated that cysteine or methionine rescued the morphogenetic defects exhibited by the met32 knockout strain, further solidifying the direct involvement of sulfur metabolism in this morphogenetic switching process. How active sulfur metabolism contributes towards fungal morphogenesis is not yet understood but given its critical requirement during this phenomenon, it would be an important question to explore and will further our overall understanding of this important biological phenomenon.
It is interesting to note that the expression of Met4, the principal transcriptional regulator of the de novo biosynthesis of sulfur-containing amino acids in S. cerevisiae along with Met32 one of its cognate DNA-binding cofactors (Blaiseau et al., 1997), are significantly downregulated when glycolysis is perturbed. Given the obligate requirement of Met4 interaction with DNA-binding proteins such as Met28, Met32, or Met31, to elicit transcriptional activation of the genes involved in the de novo biosynthesis of sulfur-containing amino acids (Blaiseau and Thomas, 1998), the observed downregulation in both Met4 and Met32 levels likely compromises the overall biosynthesis of sulfur-containing amino acids. This concurrent downregulation could explain the attenuated expression of downstream genes involved in the synthesis of methionine and cysteine. This suggests a novel regulatory mechanism wherein the expression of Met4/Met32 is coupled to glycolytic activity, under nitrogen-limiting conditions. Given that sulfur metabolism is coupled to glycolysis in C. albicans as well and contributes towards fungal morphogenesis, it is likely that this conserved regulatory axis might be critical for fungal morphogenesis in a broad range of fungal species.
Interestingly, Met4 regulation is directly dependent on Met30 (Rouillon et al., 2000). Met30 functions as a molecular switch for Met4, the master regulator of the Met gene network. Under sulfur replete conditions, Met30 ensures that Met4 remains inactive through ubiquitination and proteasomal degradation. Conversely, a reduction in sulfur availability diminishes Met30 activity, thereby de-repressing Met4 and allowing it to induce the expression of genes necessary for sulfur amino acid biosynthesis (Rouillon et al., 2000; Smothers et al., 2000; Thomas and Surdin-Kerjan, 1997). Given this regulatory mechanism, we investigated the expression of proteins involved in the sulfur assimilation pathway (Met4, Met32, Met16 and Cys3), in a viable Δmet30 deletion strain, under conditions where glycolysis was perturbed using 2DG. We observed no difference in protein levels of Met4, Met32, Met16 and Cys3, between 2DG treated and untreated cells. This indicates that, in the heterozygous deletion strain of met30 (Δmet30), Met4 is not degraded even when glycolysis is perturbed, supporting the hypothesis that perturbing active glycolysis with inhibitors like 2DG increases Met30 activity, which in turn leads to the increased degradation of Met4, which would then negatively affect Met gene transcription subsequently attenuating pseudohyphal differentiation. However, a surprising finding emerged when we examined Met30 levels in the presence of 2DG. There was no increase in the levels of Met30 in 2DG treated cells compared to untreated cells. This suggests that glycolysis influences Met30 activity through a post-translational mechanism, rather than by altering its protein abundance. Studies have shown that under cadmium stress, Met4 activity is upregulated by reducing the activity of Met30 through post-translational events, specifically by inducing the dissociation of Met30 from the core SCF complex without affecting Met30 protein levels (Yen et al., 2005). This well-established paradigm provides a compelling framework for our observation, suggesting that active glycolytic flux, similar to cadmium stress, might rapidly modulate the activity of Met30 by decreasing its association with the SCF complex to control Met4 activity, rather than through changes in protein synthesis or degradation. Future studies will be needed to delineate the precise post-translational modifications or regulatory interactions that modulate Met30 activity in response to glycolytic flux.
Intriguingly, our work extended these findings to the pathogenic fungus C. albicans, revealing a conserved role for glycolysis-mediated sulfur metabolism in hyphal differentiation, a key virulence attribute in this species (Gow et al., 2012; Moyes et al., 2016; Wilson et al., 2016). Similar to our observations in S. cerevisiae, C. albicans exhibited glycolysis-dependent hyphal formation under nitrogen limitation, which was inhibited by 2DG. Consistent with our findings in S. cerevisiae, RT-qPCR analysis revealed a significant downregulation of genes involved in de novo biosynthesis of sulfur-containing amino acids (met32, met3, met5 (ecm17), met10 and met17 (met15)) upon 2DG treatment in C. albicans. Importantly, exogenous addition of cysteine and methionine rescued the hyphal differentiation defects caused by 2DG in C. albicans, suggesting a conserved link between glycolysis and sulfur metabolism in regulating morphogenesis. Furthermore, a C. albicans strain lacking pfk1, displayed attenuated hyphal differentiation and significant downregulation of genes involved in de novo biosynthesis of sulfur-containing amino acids (met32, met3, met5 (ecm17), met10, met17 (met15)), which was rescued by cysteine and methionine supplementation, mirroring our observations in S. cerevisiae. These results clearly demonstrate that de novo biosynthesis of sulfur-containing amino acids is a critical downstream effector of glycolytic activity in promoting hyphal differentiation under nitrogen-limiting conditions, in C. albicans as well. Given that central carbon metabolic pathways like glycolysis and sulfur assimilation pathway are broadly conserved metabolic processes across a plethora of fungal species, it is possible that this novel regulatory axis might play a crucial role in modulating fungal morphogenesis in multiple fungal species and further studies are warranted to explore this. Finally, given the established link between hyphal differentiation and C. albicans virulence, we explored the impact of glycolytic perturbation on the pathogenicity of this fungus. Interestingly, RT-qPCR analysis revealed a significant downregulation of genes involved in hyphal differentiation and virulence (als3, ece1, hwp1, hyr1, ihd1, rbt1 and sap6) upon 2DG treatment and in the ΔΔpfk1 strain. Interestingly, deletions of sugar kinases or transcription factors regulating the expression of glycolysis genes or adh1 in C. albicans leads to filamentation defects and downregulation of filamentation/virulence-specific genes (Askew et al., 2009; Laurian et al., 2019; Song et al., 2019). Our data, showing transcriptional downregulation of hyphal differentiation and virulence genes following glycolytic perturbation via the deletion of pfk1, strongly corroborate these previous findings. This collectively positions glycolysis as a central metabolic nexus, rather than merely a catabolic process, profoundly influencing the pathogenic ability of C. albicans by intricately linking cellular metabolism, morphogenetic transitions, and the expression of its virulence factors. Furthermore, the ΔΔpfk1 strain displayed significantly reduced survival within murine macrophages and attenuated virulence in a murine model of systemic candidiasis, as evidenced by increased host survival, lower kidney fungal burden, and reduced tissue colonization. Remarkably, exogenous sulfur supplementation (N-acetyl cysteine (NAC)) to mice rescued the attenuated virulence exhibited by ΔΔpfk1. Specifically, mice receiving NAC-containing distilled water and infected with ΔΔpfk1 showed significantly reduced survival and increased tissue colonization compared to those given normal distilled water. These findings underscore the critical role of glycolysis, and by extension, glycolysis-dependent sulfur metabolism, in the virulence of C. albicans.
In conclusion, our study unveils a conserved metabolic network that orchestrates fungal morphogenesis in response to nutrient availability. We demonstrate that active glycolysis, fueled by fermentable carbon sources in a cAMP-PKA independent manner, plays a crucial role in regulating the de novo biosynthesis of sulfur-containing amino acids in Met30-dependent manner, which in turn are essential for the morphological transitions in both S. cerevisiae and C. albicans (Fig. 6). The attenuation of C. albicans virulence upon disruption of this pathway highlights the potential of targeting specific fungal metabolic networks as a novel antifungal strategy. Future research in our laboratory will focus on elucidating the precise molecular mechanisms by which glycolytic intermediates or downstream signals regulate the expression of sulfur metabolism pathway and how sulfur-containing amino acids contribute to the cellular processes underlying fungal morphogenesis.

Glycolysis-dependent Sulfur Metabolism Orchestrates Morphological Transitions and Virulence in Fungi.
Active glycolysis plays a key regulatory role in the de novo biosynthesis of sulfur-containing amino acids by modulating the activity of SCFMet30 complex (SCFMet30 complex includes Met30, Skp1, Cdc53, Hrt1 and Cdc34) which in turn affects the expression of genes involved in this process (In S. cerevisiae, Met transcription complex includes Met4, Met28, Cbf1, Met31, and Met32, whereas in C. albicans, the characterized components of this complex include Met4, Cbf1, and Met32. Hence, Met28 and Met31, unique to the Met transcription complex of S. cerevisiae are represented with dotted borders). Glycolysis-dependent sulfur metabolism in turn, is critical for fungal morphogenesis in both species and the ability of C. albicans to survive within macrophages and cause systemic infection in a murine model of candidiasis. Perturbation of active glycolysis increases Met30 activity, which leads to the increased degradation of Met4, resulting in the reduced expression of genes involved in the de novo biosynthesis of sulfur-containing amino acids. This in turn attenuates fungal morphogenesis in both species and the ability of C. albicans to survive within macrophages and cause systemic infection in a murine model of candidiasis. This figure was created using BioRender.com.
Materials and Methods
Yeast Strains
The prototrophic diploid strains, ∑1278b or CEN.PK were used in all the experiments involving S. cerevisiae. The prototrophic diploid strain, SC5314 was exclusively used in all the experiments involving C. albicans. The yeast strains used in this study are listed in Suppl. Table 1.
Generation of S. cerevisiae Gene Knockout and Epitope-tagged Strains
A standard PCR-based technique was used to amplify resistance cassettes (G418, HygB, or NAT) with flanking sequences in order to carry out gene deletions. The target gene was then replaced by homologous recombination using the lithium acetate-based transformation (Schiestl and Gietz, 1989). Similarly, C-terminal epitope-tagged yeast strains were generated using a PCR-based technique. This involved amplifying resistance cassettes, flanked by sequences homologous to the C-terminal region of the target gene, along with the desired epitope tags. The targeted gene was then tagged at its C-terminus through homologous recombination, using the lithium acetate-based transformation method. Primers used for these experiments are listed in Suppl.Table 2.
Generation of C. albicans Gene Knockout Strains
The C. albicans strains used in this study are listed in Suppl.Table 1. The previously established SAT1-flipper method was used with some modifications, for the deletion of the desired genes (Reuß et al., 2004). To generate the pfk1 deletion (ΔΔpfk1) strain, a DNA fragment carrying the flanking regions of pfk1 (100 bps homology) and the SAT1-flipper cassette was transformed into C. albicans wild-type SC5314, using electroporation. Transformants were selected based on their nourseothricin resistance. PCR confirmation was used to validate the transformants. The SAT1-flipper cassette was then excised from the pfk1 locus by growing the cells in YPM medium (10g/l yeast extract, 20g/l peptone, and 2% (w/v) maltose) to induce expression of the MAL2 promoter-regulated recombinase. To knock out the second allele of pfk1, the heterozygous pfk1 deletion mutants (Δpfk1/PFK1) were used for the transformation wherein the DNA fragment carrying the flanking regions of pfk1 (100 bps homology) and the SAT1-flipper cassette was transformed into C. albicans Δpfk1/PFK1 SC5314. Transformants were selected based on their nourseothricin resistance. PCR confirmation was used to validate the transformants. Primers used to generate or confirm the deletion strains are listed in Suppl.Table 2.
Media and Growth Conditions
YPD broth containing 10g/l yeast extract, 20g/l peptone and 2% (w/v) glucose was used for the overnight growth of the cultures. Nitrogen-limiting medium containing 1.7g/l yeast nitrogen base without amino acids or ammonium sulfate, 2% (w/v) glucose (SLAD) or various different fermentable or non-fermentable carbon sources, 50µM ammonium sulfate and 2% (w/v) agar which was extensively washed, to remove excess nitrogen (Gimeno et al., 1992) was used for all experiments involving pseudohyphal or hyphal differentiation. Overnight cultures of various strains were grown in YPD broth and incubated at 30 °C with continuous shaking at 200 rotations per minute (RPM). Reagents and tools used in this study are mentioned in Suppl.Table 3.
Pseudohyphal Differentiation Assay
To induce pseudohyphal differentiation in S. cerevisiae, overnight cultures of various strains (listed in Suppl.Table 1) were spotted on nitrogen-limiting medium (SLA) containing 2% glucose (SLAD). In order to check the effect of glycolysis inhibitors, 0.05% concentration of 2-Deoxy-D-Glucose (2DG) and 0.2% concentration of sodium citrate (NaCi), on pseudohyphal differentiation, wild-type was spotted on SLAD containing sub-inhibitory concentrations of 2DG (0.05%) or NaCi (0.2%). Plates were incubated for 10 days at 30 °C (Chandarlapaty and Errede, 1998; González et al., 2017). After 10 days, imaging was done using Olympus MVX10 stereo microscope. Cells from colonies were isolated and single cells were imaged using Zeiss Apotome microscope and the percentage of pseudohyphal cells was calculated using ImageJ. Cells with a length/width ratio of 2 or more were considered as pseudohyphal cells as described previously (Schröder et al., 2000). Statistical analysis was done using unpaired t-test or one-way ANOVA test. Error bars represent SEM. All the experiments were performed independently, thrice.
Hyphal Differentiation Assay
To induce hyphal differentiation in C. albicans, overnight cultures of each strain (listed in Suppl.Table 1) were spotted on nitrogen-limiting medium (SLAD). To check the effect of glycolysis inhibitor, 2-Deoxy-D-Glucose (2DG) on hyphal differentiation, wild-type was spotted on SLAD containing sub-inhibitory concentration of 2DG (0.2%). All the plates were incubated for 7 days at 37 °C (Sánchez-Martínez and Pérez-Martín, 2002; Song et al., 2019). After 7 days, cells from colonies were isolated and single cells were imaged using Zeiss Apotome microscope and percentage of hyphal cells was calculated using ImageJ. Cells with a length/width ratio of 4.5 or more were considered as hyphal cells as described previously (Su et al., 2018). Statistical analysis was done using unpaired t-test. Error bars represent SEM. All the experiments were performed independently, thrice.
Growth Curves
Growth curve analysis in the presence of glycolysis inhibitors (2DG or NaCi) was performed as follows. Overnight grown cultures of wild-type strains (∑1278b or CEN.PK or SC5314) were diluted to OD600=0.01 in fresh SLAD medium with and without 2DG or NaCi and allowed to grow at 30 °C for 24 hours. OD600 was recorded at 3-h intervals. Growth curve analysis using various deletion strains were performed as follows. Overnight grown cultures of various deletion strains along with wild-type strains (∑1278b or CEN.PK or SC5314) were diluted to OD600=0.01 in fresh SLAD and allowed to grow at 30 °C for 24 hours. OD600 was recorded at 3-h intervals. Graphs were prepared using GraphPad Prism. Error bars represent SEM. All the experiments were performed independently, thrice.
cAMP Add-back Assays
In order to perform cAMP add-back assays, overnight cultures of wild-type ∑1278b and various deletion strains of S. cerevisiae (listed in Suppl.Table 1) were spotted on SLAD or SLAD containing sub-inhibitory concentration of 2DG or NaCi, in the presence and absence of cAMP (1mM) (Lorenz and Heitman, 1997). Plates were incubated for 10 days at 30 °C to induce pseudohyphal differentiation in S. cerevisiae. After 10 days, imaging was done using Olympus MVX10 stereo microscope. All the experiments were performed independently, thrice.
Sulfur Add-back Assays
In order to perform sulfur add-back assays, overnight cultures of various strains of S. cerevisiae or C. albicans (listed in Suppl.Table 1) were spotted on SLAD or SLAD containing sub-inhibitory concentration of 2DG, in the presence and absence of sulfur-containing compounds including cysteine, methionine. Plates were incubated for 10 days at 30 °C to induce pseudohyphal differentiation in S. cerevisiae. After 10 days, imaging was done using Olympus MVX10 stereo microscope. Cells from colonies were isolated and single cells were imaged using Zeiss Apotome microscope and percentage of pseudohyphal cells were calculated using ImageJ. For hyphal differentiation assays in C. albicans, plates were incubated for 7 days at 37 °C. After 7 days, cells from colonies were isolated and single cells were imaged using Zeiss Apotome microscope and percentage of hyphal cells were calculated using ImageJ. Statistical analysis was done using unpaired t-test or one-way ANOVA test. Error bars represent SEM. All the experiments were performed independently, thrice.
Whole genome sequencing
Sample Collection
Strains of S. cerevisiae (ΔΔpfk1 or ΔΔadh1 along with wild-type strain ∑1278b) or C. albicans (ΔΔpfk1 along with wild-type strain SC5314) were grown in YPD medium at 30 °C for overnight.
Genomic DNA Extraction
Genomic DNA from the samples was extracted using phenol-chloroform method. Following RNAse treatment to eliminate any contaminating RNA from the sample, the extracted DNA was submitted to the CCMB next-generation sequencing facility for whole genome sequencing.
Whole genome sequencing Analysis
Raw sequencing reads were obtained in FASTQ format. The reference genome and annotation files for the S. cerevisiae strain ∑1278b were downloaded from the Saccharomyces Genome Database (SGD) and reference genome and annotation files for the C, albicans strain SC5314 were downloaded from the NCBI (National Center for Biotechnology Information). Quality assessment of the raw reads was performed using FastQC (version 0.12.1). Adapter sequences were trimmed using Cutadapt (version 4.6). The reference genome was indexed with Hisat2 (version 2.2.1), and Samtools (version 1.20) was used to filter out multi-mapped reads and convert SAM files to BAM format. Visualisation of reads was done using IGV software (version 2.16.2).
RNA-Sequencing
Sample Collection
Overnight culture of S. cereviae wild-type ∑1278b was spotted on nitrogen-limiting media containing 2% (w/v) glucose (SLAD) with and without sub-inhibitory concentration of 2-Deoxy-D-Glucose (2DG). The plates were incubated at 30 °C and colonies were isolated on day 5 (D-5) and day 10 (D-10).
RNA-Extraction
Total RNA from the sample was extracted using the protocol mentioned in the Ribopure RNA purification kit (yeast) (cat. #AM1926). Following a DNAse treatment to eliminate any genomic DNA from the sample, the extracted RNA was submitted to the CCMB next-generation sequencing facility for RNA-Seqencing. The NovaSeq 6000 equipment was used to sequence transcriptomes.
RNA-Sequencing Analysis
Raw sequencing reads were obtained in FASTQ format. The reference genome and annotation files for the S. cerevisiae strain ∑1278b were downloaded from the Saccharomyces Genome Database (SGD). Quality assessment of the raw reads was performed using FastQC (version 0.12.1). Adapter sequences were trimmed using Cutadapt (version 4.6). The reference genome was indexed with Hisat2 (version 2.2.1), and Samtools (version 1.20) was used to filter out multi-mapped reads and convert SAM files to BAM format. Read counting for uniquely mapped reads was done with FeatureCounts, using gene annotation data from the reference S. cerevisiae ∑1278b genome. Normalization and differential gene expression analysis were performed using R (version 4.3.2). Genes with a log2 fold change of ≥1 were considered to be upregulated, indicating at least a doubling in expression under the given condition, while those with a log2 fold change of ≤ −1 were considered to be downregulated.
Protein Isolation
For Western blotting experiments, in order to isolate proteins from liquid cultures, overnight yeast cultures were back-diluted in 10 ml of liquid SLAD in the presence and absence of 2-Deoxy-D-Glucose (2DG). The yeast cells were grown till an OD600 of ∼1 and cells were harvested by centrifugation at 3000 rpm for 10 minutes at room temperature (RT). To isolate proteins from colonies, overnight cultures of pertinent strains were spotted on SLAD with and without sub-inhibitory concentration of 2-Deoxy-D-Glucose (2DG) (0.05%) and colonies were isolated and resuspended in 1ml of 1X PBS (Phosphate Buffered Saline) after 5 days (D-5). Cells were harvested by centrifugation at 14000 rpm for 10 minutes at room temperature. Pelleted cells (from liquid cultures and colonies) were resuspended in 300µl of ice cold 10% (w/v) TCA solution and disrupted using bead-beating by addition of 150µl of glass beads at 4°C. Supernatant was transferred to another tube. Next, proteins were pelleted by centrifugation at 14,000 rpm for 10 minutes at 4°C and supernatant was discarded. Each pellet was resuspended in 400µl of SDS/glycerol buffer (7.3% (w/v) SDS, 29% (v/v) glycerol and 83.3mM tris-base) and boiled at 95°C for 10 minutes with occasional vortexing followed by centrifugation at 14,000 rpm for 10 minutes at room temperature. Supernatant was transferred to a fresh tube and used for protein estimation using BCA protein estimation kit (cat. #23227). Finally, all samples were diluted in SDS-PAGE loading buffer (100mM Tris-Cl, 4% (w/v) SDS, 0.02% (w/v) bromophenol blue, 20% (w/v), glycerol and 7.5% (v/v) β-mercaptoethanol) and heated for 5 minutes at 90 °C.
Immunodetection
Protein samples were separated using SDS-PAGE on polyacrylamide gels and transferred to PVDF membranes with a Mini Trans-Blot Cell module (Bio-Rad). The membranes were blocked for 1 hour at room temperature with 5% (w/v) non-fat dry milk in 1X TBST buffer (20 mM Tris-Base, pH 7.6, 150 mM NaCl, 0.1% (v/v) Tween-20) with gentle orbital shaking. After blocking, the membranes were incubated overnight at 4°C with primary antibodies (1:2000 dilution) in 1X TBST buffer containing 5% (w/v) non-fat dry milk. Following this, the membranes were washed three times for 10 minutes in 1X TBST and then incubated for 1 hour at room temperature with HRP-conjugated secondary antibodies (1:4000 dilution) in 1X TBST with 5% (w/v) non-fat dry milk. After three additional washes in 1X TBST, chemiluminescent substrate was added and bands were detected using the SuperSignal West chemiluminescent substrate (cat. #34579, cat. #34094) and visualized using the Bio-Rad ChemiDocTM MP imaging system. All the experiments were performed independently, thrice.
Antibodies
HA-Tag (C29F4) Rabbit mAb (cat. #3724), β-actin (8H10D10) Mouse mAb (cat. #3700), Goat anti-Rabbit IgG-HRP conjugate (cat. #7074) and Horse anti-Mouse IgG-HRP conjugate (cat. #7076) were purchased from Cell Signaling Technology. Pgk1 Mouse mAb (cat. #sc-130335) was purchased from Santa Cruz Biotechnology.
Densitometric Analysis
Raw images of Western blots were analysed using ImageJ to normalize target protein expression with the expression of housekeeping proteins including PGK1 and β-actin in order to make densitometric graphs. Statistical analysis was done using unpaired t-test, with error bars representing SEM. All the experiments were performed independently, thrice.
Quantitative Real Time Polymerase Chain Reaction (RT-qPCR) Assay
Overnight cultures of C. albicans wild-type SC5314 was spotted on SLAD with and without sub-inhibitory concentration of 2-Deoxy-D-Glucose (2DG) (0.2%) or C. albicans ΔΔpfk1 strain along with wild-type SC5314 was spotted on SLAD and plates were incubated at 37 °C for 4 days. After 4 days, colonies were isolated and total RNA was extracted using the protocol mentioned in the Ribopure RNA purification kit, yeast (cat. #AM1926). cDNA synthesis was carried out using PrimeScript cDNA synthesis kit (cat. #6110A). Primers used for RT-qPCR were designed using PrimeQuest tool of Integrated DNA Technologies (IDT). RT-qPCR was performed using iTaq Universal SYBR green mastermix (cat. #1725121). Actin was used as housekeeping gene for normalization. The results were analysed using the 2-ΔΔCT method (Winer et al., 1999) and statistical analysis was done using one-way ANOVA test, with error bars representing SEM. All the experiments were performed independently, thrice. Primers used in the RT-qPCR experiments are mentioned in Suppl.Table 2.
In vitro Microbial Survival Assay
In vitro microbial survival assay in macrophages was conducted as described in previous studies with some modifications (Netea et al., 2002; Pradhan et al., 2016). RAW 264.7 macrophages (1×10⁶ cells) were cultured in six-well plates with DMEM supplemented with 10% FBS. After 16 hours of incubation, the cells were treated with C. albicans wild-type SC5314 or ΔΔpfk1 at a multiplicity of infection (MOI) of 10, for 1 hour in antibiotic-free DMEM containing 10% FBS. The media was then removed, and the cells were washed six times with 1X PBS to eliminate any free, non-phagocytosed fungal cells. Next, the cells were lysed with 350 µl of lysis buffer containing 0.025% SDS in 1X PBS. The lysate volume was adjusted to 1 ml with 1X PBS, and a 100 µl undiluted aliquot was plated on YPD agar plate. Serial dilutions of the lysate were also plated on YPD agar plates. These plates were incubated overnight at 37 °C. The following day, colony forming units (CFU) were counted and plotted. The percentage of survival was calculated as the number of CFU in the presence of macrophages divided by the number of CFU in the absence of macrophages. Statistical analysis was done using unpaired t-test, with error bars representing SEM. All the experiments were performed independently, thrice.
In vivo Murine Model of Systemic Candidiasis
In vivo infection experiments were performed as previously described (Majer et al., 2012). Briefly, C. albicans wild-type SC5314 and ΔΔpfk1 were grown overnight in YPD medium at 30 °C. The overnight cultures were back-diluted in 10 ml of YPD liquid medium and allowed to grow till and OD of ∼0.6, then washed and resuspended (1 × 10⁷ CFU) in 1X PBS. C57BL/6 wild-type mice, aged 6 to 8 weeks, were injected intravenously via lateral tail vein with C. albicans wild-type SC5314 (1 × 107 CFU) or ΔΔpfk1 (1 × 107 CFU). After infection, mice were monitored for survival and daily for signs of morbidity and mortality were recorded over a period of 21 days. Survival data were analysed using the Log-rank Mantel-Cox test In order to perform in vivo sulfur supplementation, N-acetyl cysteine (NAC) was dissolved in distilled water at a concentration of 6mg/ml and administrated orally to mice before 72 hours of C. albicans infection. Normal distilled water administration to mice was used as control. After 72 hours of NAC administration, C57BL/6 wild-type mice, aged 6 to 8 weeks, were infected with SC5314 (1 × 107 CFU) or ΔΔpfk1 strain (1 × 107 CFU), intravenously via lateral tail vein injection. After infection, mice were monitored for survival and daily for signs of morbidity and mortality were recorded over a period of 21 days. Survival data were analysed using the Log-rank Mantel-Cox test.
Determination of Fungal Burden
To assess fungal burden, mice injected with C. albicans wild-type SC5314 or ΔΔpfk1 were sacrificed at ∼2 days and kidneys were harvested from infected mice. Kidneys were weighed and then homogenized using a sterilized and PBS-washed Dounce homogenizer (cat. #D8938). Kidney suspensions were then diluted in 1 ml of 1X PBS and serially diluted. Then 100µl of dilutions was plated on YPD agar plates and allowed to grow at 37 °C. Colony forming units (CFU) were calculated by the following formula: (number of colonies × dilution factor × 10)/organ weight in grams (g). Statistical analysis was done using unpaired t-test. Error bars represent SEM. All the experiments were performed independently, thrice.
Histopathology
For histopathological examination, kidneys were collected and fixed in 10% neutral buffered formalin for about two weeks. After fixation, the kidneys were processed, embedded in paraffin, and sectioned to 4 µm thickness. The tissue sections were then stained using the GMS stain (Grocott Methenamine Silver stain) and the protocol mentioned in the GMS staining kit manual (cat. #AB287884) was followed. All the experiments were performed independently, thrice.
Illustrations
Figure illustrations were created using Biorender (https://app.biorender.com).
Animal Ethics Statement
All animal experiments in this manuscript were reviewed and approved by the Institutional Animal Ethics Committee (IAEC) of CSIR-Centre for Cellular and Molecular Biology (Approval number: 52/2024-C).
Supporting Information

Confirmation of pfk1 and adh1 Deletion in S. cerevisiae.
(A) Confirmation of ΔΔpfk1 strain using whole genome sequencing. IGV software (version 2.16.2) was used for data visualisation. Red box represents gene locus of pfk1 in wild-type (∑1278b) and ΔΔpfk1 strains. (B) Confirmation of ΔΔadh1 strain using whole genome sequencing. IGV software (version 2.16.2) was used for data visualisation. Red box represents gene locus of adh1 in wild-type (∑1278b) and ΔΔadh1 strains. This figure was created using BioRender.com.

Comparative Transcriptomics Identifies the Role of Glycolysis During Pseudohypal Differentiation in S. cerevisiae.
(A) Wild-type ∑1278b was spotted on SLAD, incubated for 10 days at 30 °C. Colonies were imaged every day, for 10 consecutive days (D-1 to D-10), using Olympus MVX10 stereo microscope. Scale bar represents 1mm for whole colony images. (B) Heatmaps represent the differentially expressed genes involved in amino acid biosynthesis and transporters, in 2DG treated cells compared to untreated cells isolated from D-5 and D-10 colonies, respectively (n=3). Scale bar represents Z-Score. This figure was created using BioRender.com.

Expression of Proteins Involved in the de novo Biosynthesis of Sulfur-containing Amino Acids in Response to Glycolysis Perturbation.
(A) Epitope-tagged strains of various proteins involved in the de novo biosynthesis of sulfur-containing amino acids including Met4, Met32, Met16, Met10, Cys4 and Cys3 were spotted on SLAD in the presence and absence of sub-inhibitory concentration of 2DG following which, cells from colonies were isolated at day 5 and levels of these proteins were assessed using Western blotting. Raw images of Western blots were analysed using ImageJ to normalize targeted protein expression with the expression of housekeeping protein-β-actin, in order to generate densitometric graphs. Statistical analysis was done using unpaired t-test, **(P<0.01) and *(P<0.05). Error bars represent SEM. (B) Epitope-tagged strains of various proteins involved in the de novo biosynthesis of sulfur-containing amino acids including Met32, Met16, Met10 and Cys3 were grown in liquid SLAD and then treated with sub-inhibitory concentration of 2DG for 24 hours after which protein expression was checked using Western blotting (n=3). Pgk1 was used as loading control. (C) Raw images of Western blots were analysed using ImageJ to normalize targeted protein expression with the expression of housekeeping protein-Pgk1, in order to generate densitometric graphs. Statistical analysis was done using unpaired t-test, ***(P<0.001), **(P<0.01) and *(P<0.05). Error bars represent SEM. This figure was created using BioRender.com.

Monitoring of Overall Growth Rate of ΔΔmet32 Strain.
(A) Growth curve analysis was performed to monitor overall growth of wild-type and ΔΔmet32 knockout strain on SLAD. Overnight grown cultures of deletion strain (ΔΔmet32) along with wild-type strain (∑1278b) were diluted to OD600=0.01 in fresh SLAD and allowed to grow at 30 °C for 24 hours. OD600 was recorded at 3-h intervals. This figure was created using BioRender.com.

Confirmation of pfk1 Deletion in C. albicans.
(A) Confirmation of ΔΔpfk1 strain using whole genome sequencing. IGV software (version 2.16.2) was used for data visualisation. Red box represents gene locus of pfk1 in wild-type (SC5314) and ΔΔpfk1 strains. This figure was created using BioRender.com.

Yeast Strains Used in this Study



Oligonucleotides Used in this Study



List of Reagents and Tools Used in this Study
Acknowledgements
We gratefully acknowledge the invaluable support and resources offered by CSIR-Centre for Cellular and Molecular Biology (CCMB) central facilities including Advanced Microscopy and Imaging facility (AMIF), Next Generation Sequencing (NGS) facility and Animal House facility. We acknowledge help from T. Avinash Raj (CSIR-CCMB) for preparation of samples for histological analysis. DS thanks University Grants Commission (UGC), India for their research fellowship. SV acknowledges funding from Indian Council of Medical Research (ICMR) (IIRPSG-2024-01-02717), India; Anusandhan National Research Foundation (ANRF) (SRG/2023/000470), India; Council of Scientific & Industrial Research (CSIR) (FTT070505), India; and DBT-Wellcome Trust India Alliance (IA/E/16/1/502996), India.
Additional information
Data Availability
All relevant data are within the paper and its supporting information files, except for RNA-Seq data. RNA-Seq data are deposited in NCBI’s SRA database and are accessible through BioProject accession number PRJNA1263201. (https://www.ncbi.nlm.nih.gov/bioproject/PRJNA1263201)
Author Contributions
Dhrumi Shah: Conceptualization; Investigation; Formal analysis; Visualization; Writing-original draft. Nikita Rewatkar: Investigation. Adishree M: Investigation. Siddhi Gupta: Investigation. Sudharsan Mathivathanan: Investigation. Sayantani Biswas: Investigation. Sriram Varahan: Conceptualization; Supervision; Funding acquisition; Investigation; Formal analysis; Visualization; Writing-original draft; Writing-review and editing.
Funding
Indian Council of Medical Research (IIRPSG-2024-01-02717)
Anusandhan National Research Foundation (SRG/2023/000470)
Council of Scientific and Industrial Research (FTT070505)
DBT-Wellcome Trust India Alliance (IA/E/16/1/502996)
References
- Transcriptional regulation of carbohydrate metabolism in the human pathogen Candida albicansPLoS pathogens 5:e1000612https://doi.org/10.1371/journal.ppat.1000612Google Scholar
- The Candida albicans HYR1 gene, which is activated in response to hyphal development, belongs to a gene family encoding yeast cell wall proteinsJournal of bacteriology 178:5353–5360https://doi.org/10.1128/jb.178.18.5353-5360.1996Google Scholar
- An examination of the crabtree effect in Saccharomyces cerevisiae: The role of respiratory adaptationJ Gen Microbiol https://doi.org/10.1099/00221287-114-2-267Google Scholar
- The primary structure of the Saccharomyces cerevisiae gene for alcohol dehydrogenaseThe Journal of biological chemistry 257:3018–3025Google Scholar
- Environmental sensing and signal transduction pathways regulating morphopathogenic determinants of Candida albicansMicrobiology and molecular biology reviews: MMBR 71:348–376https://doi.org/10.1128/MMBR.00009-06Google Scholar
- Met31p and Met32p, two related zinc finger proteins, are involved in transcriptional regulation of yeast sulfur amino acid metabolismMolecular and cellular biology 17:3640–3648https://doi.org/10.1128/MCB.17.7.3640Google Scholar
- Multiple transcriptional activation complexes tether the yeast activator Met4 to DNAThe EMBO journal 17:6327–6336https://doi.org/10.1093/emboj/17.21.6327Google Scholar
- Nutritional control of growth and development in yeastGenetics 192:73–105https://doi.org/10.1534/genetics.111.135731Google Scholar
- Ash1, a daughter cell-specific protein, is required for pseudohyphal growth of Saccharomyces cerevisiaeMolecular and cellular biology 18:2884–2891https://doi.org/10.1128/MCB.18.5.2884Google Scholar
- Adaptation of Candida albicans to Reactive Sulfur SpeciesGenetics 206:151–162https://doi.org/10.1534/genetics.116.199679Google Scholar
- A View of Fungal EcologyMycologia 81:1–19https://doi.org/10.1080/00275514.1989.12025620Google Scholar
- 2-Desoxy-D-glucose as an antagonist of glucose in yeast fermentation, Journal of the Franklin InstituteVolume 253:354–360https://doi.org/10.1016/0016-0032(52)90852-1Google Scholar
- Roles of the Candida albicans mitogen-activated protein kinase homolog, Cek1p, in hyphal development and systemic candidiasisInfection and immunity 66:2713–2721https://doi.org/10.1128/IAI.66.6.2713-2721.1998Google Scholar
- The Crabtree effect: a regulatory system in yeastJournal of general microbiology 44:149–156https://doi.org/10.1099/00221287-44-2-149Google Scholar
- Virulence Traits Contributing to Pathogenicity of Candida SpeciesJ Microbiol Exp 5:00140https://doi.org/10.15406/jmen.2017.05.00140Google Scholar
- The catabolism of amino acids to long chain and complex alcohols in Saccharomyces cerevisiaeThe Journal of biological chemistry 278:8028–8034https://doi.org/10.1074/jbc.M211914200Google Scholar
- How yeast re-programmes its transcriptional profile in response to different nutrient impulsesBMC systems biology 5:148https://doi.org/10.1186/1752-0509-5-148Google Scholar
- Partial purification and properties of pyruvate kinase and its regulatory role during lipid accumulation by the oleaginous yeast Rhodosporidium toruloides CBS 14Canadian Journal of Microbiology 31:479–484https://doi.org/10.1139/m85-089Google Scholar
- Nutrient acquisition by pathogenic fungi: nutrient availability, pathway regulation, and differences in substrate utilizationInternational journal of medical microbiology: IJMM 301:400–407https://doi.org/10.1016/j.ijmm.2011.04.007Google Scholar
- Biosynthesis and regulation of fructose-1,6-bisphosphatase and phosphofructokinase in Saccharomyces cerevisiae grown in the presence of glucose and gluconeogenic carbon sourcesJournal of bacteriology 136:647–656https://doi.org/10.1128/jb.136.2.647-656.1978Google Scholar
- Control of pseudohyphae formation in Saccharomyces cerevisiaeFEMS microbiology reviews 25:107–123https://doi.org/10.1111/j.1574-6976.2001.tb00573.xGoogle Scholar
- Unipolar cell divisions in the yeast S. cerevisiae lead to filamentous growth: regulation by starvation and RASCell 68:1077–1090https://doi.org/10.1016/0092-8674(92)90079-rGoogle Scholar
- Role of Mitochondrial Retrograde Pathway in Regulating Ethanol-Inducible Filamentous Growth in YeastFrontiers in physiology 8:148https://doi.org/10.3389/fphys.2017.00148Google Scholar
- Fungal morphogenesis and host invasionCurrent opinion in microbiology 5:366–371https://doi.org/10.1016/s1369-5274(02)00338-7Google Scholar
- Candida albicans morphogenesis and host defence: discriminating invasion from colonizationNature reviews. Microbiology 10:112–122https://doi.org/10.1038/nrmicro2711Google Scholar
- The Galpha protein Gpa2 controls yeast differentiation by interacting with kelch repeat proteins that mimic Gbeta subunitsMolecular cell 10:163–173https://doi.org/10.1016/s1097-2765(02)00569-5Google Scholar
- Construction and physiological characterization of mutants disrupted in the phosphofructokinase genes of Saccharomyces cerevisiaeCurrent genetics 11:227–234https://doi.org/10.1007/BF00420611Google Scholar
- Virulence assessment of six major pathogenic Candida species in the mouse model of invasive candidiasis caused by fungal translocationScientific reports 10:3814https://doi.org/10.1038/s41598-020-60792-yGoogle Scholar
- Regulation of phenotypic transitions in the fungal pathogen Candida albicansVirulence 3:251–261https://doi.org/10.4161/viru.20010Google Scholar
- Fungal immune evasion in a model host-pathogen interaction: Candida albicans versus macrophagesPLoS pathogens 9:e1003741https://doi.org/10.1371/journal.ppat.1003741Google Scholar
- Cdc34 and the F-box protein Met30 are required for degradation of the Cdk-inhibitory kinase Swe1Genes & development 12:2587–2597https://doi.org/10.1101/gad.12.16.2587Google Scholar
- A Saccharomyces cerevisiae G-protein coupled receptor, Gpr1, is specifically required for glucose activation of the cAMP pathway during the transition to growth on glucoseMolecular microbiology 32:1002–1012https://doi.org/10.1046/j.1365-2958.1999.01413.xGoogle Scholar
- The Complex Genetic Basis and Multilayered Regulatory Control of Yeast Pseudohyphal GrowthAnnual review of genetics 55:1–21https://doi.org/10.1146/annurev-genet-071719-020249Google Scholar
- Hexokinase and Glucokinases Are Essential for Fitness and Virulence in the Pathogenic Yeast Candida albicansFrontiers in microbiology 10:327https://doi.org/10.3389/fmicb.2019.00327Google Scholar
- Multiple TORC1-associated proteins regulate nitrogen starvation-dependent cellular differentiation in Saccharomyces cerevisiaePloS one 6:e26081https://doi.org/10.1371/journal.pone.0026081Google Scholar
- ECM17-dependent methionine/cysteine biosynthesis contributes to biofilm formation in Candida albicansFungal genetics and biology: FG & B 51:50–59https://doi.org/10.1016/j.fgb.2012.11.010Google Scholar
- Fungal morphogenesisCold Spring Harbor perspectives in medicine 5:a019679https://doi.org/10.1101/cshperspect.a019679Google Scholar
- Organ-specific innate immune responses in a mouse model of invasive candidiasisJournal of innate immunity 3:180–199https://doi.org/10.1159/000321157Google Scholar
- Candida albicans Als3, a multifunctional adhesin and invasinEukaryotic cell 10:168–173https://doi.org/10.1128/EC.00279-10Google Scholar
- Phosphofructokinase mutants of yeast. Biochemistry and geneticsThe Journal of biological chemistry 258:1444–1449Google Scholar
- Yeast pseudohyphal growth is regulated by GPA2, a G protein alpha homologThe EMBO journal 16:7008–7018https://doi.org/10.1093/emboj/16.23.7008Google Scholar
- Regulators of pseudohyphal differentiation in Saccharomyces cerevisiae identified through multicopy suppressor analysis in ammonium permease mutant strainsGenetics 150:1443–1457https://doi.org/10.1093/genetics/150.4.1443Google Scholar
- The G protein-coupled receptor gpr1 is a nutrient sensor that regulates pseudohyphal differentiation in Saccharomyces cerevisiaeGenetics 154:609–622https://doi.org/10.1093/genetics/154.2.609Google Scholar
- The G protein-coupled receptor Gpr1 and the Galpha protein Gpa2 act through the cAMP-protein kinase A pathway to induce morphogenesis in Candida albicansMolecular biology of the cell 16:1971–1986https://doi.org/10.1091/mbc.e04-09-0780Google Scholar
- Type I interferons promote fatal immunopathology by regulating inflammatory monocytes and neutrophils during Candida infectionsPLoS pathogens 8:e1002811https://doi.org/10.1371/journal.ppat.1002811Google Scholar
- Changes in yeast amino acid pool with respiratory versus fermentative metabolismBiotechnology and bioengineering 40:643–649https://doi.org/10.1002/bit.260400602Google Scholar
- A core filamentation response network in Candida albicans is restricted to eight genesPloS one 8:e58613https://doi.org/10.1371/journal.pone.0058613Google Scholar
- In Fungal Intracellular Pathogenesis, Form Determines FatemBio 9:e02092–18https://doi.org/10.1128/mBio.02092-18Google Scholar
- Fungal Pathogens: Shape-Shifting InvadersTrends in microbiology 28:922–933https://doi.org/10.1016/j.tim.2020.05.001Google Scholar
- Rbt1 protein domains analysis in Candida albicans brings insights into hyphal surface modifications and Rbt1 potential role during adhesion and biofilm formationPloS one 8:e82395https://doi.org/10.1371/journal.pone.0082395Google Scholar
- Candidalysin is a fungal peptide toxin critical for mucosal infectionNature 532:64–68https://doi.org/10.1038/nature17625Google Scholar
- Transcription profiling of Candida albicans cells undergoing the yeast-to-hyphal transitionMolecular biology of the cell 13:3452–3465https://doi.org/10.1091/mbc.e02-05-0272Google Scholar
- Fungal evolution: major ecological adaptations and evolutionary transitionsBiological reviews of the Cambridge Philosophical Society 94:1443–1476https://doi.org/10.1111/brv.12510Google Scholar
- The role of toll-like receptor (TLR) 2 and TLR4 in the host defense against disseminated candidiasisThe Journal of infectious diseases 185:1483–1489https://doi.org/10.1086/340511Google Scholar
- Glucose regulation of Saccharomyces cerevisiae cell cycle genesEukaryotic cell 2:143–149https://doi.org/10.1128/EC.2.1.143-149.2003Google Scholar
- Life within a community: benefit to yeast long-term survivalFEMS microbiology reviews 30:806–824https://doi.org/10.1111/j.1574-6976.2006.00034.xGoogle Scholar
- Signal transduction cascades regulating pseudohyphal differentiation of Saccharomyces cerevisiaeCurrent opinion in microbiology 3:567–572https://doi.org/10.1016/s1369-5274(00)00142-9Google Scholar
- Invasive candidiasisNature reviews. Disease primers 4:18026https://doi.org/10.1038/nrdp.2018.26Google Scholar
- Fructose-1,6-bisphosphate couples glycolytic flux to activation of RasNature communications 8:922https://doi.org/10.1038/s41467-017-01019-zGoogle Scholar
- Comparative Analysis of the Effects of Two Probiotic Bacterial Strains on Metabolism and Innate Immunity in the RAW 264.7 Murine Macrophage Cell LineProbiotics and antimicrobial proteins 8:73–84https://doi.org/10.1007/s12602-016-9211-4Google Scholar
- Advancements and challenges in antifungal therapeutic developmentClinical microbiology reviews 37:e0014223https://doi.org/10.1128/cmr.00142-23Google Scholar
- The SAT1 flipper, an optimized tool for gene disruption in Candida albicansGene 341:119–127https://doi.org/10.1016/j.gene.2004.06.021Google Scholar
- Processing of Candida albicans Ece1p Is Critical for Candidalysin Maturation and Fungal VirulencemBio 9:e02178–17https://doi.org/10.1128/mBio.02178-17Google Scholar
- Fungal Morphogenesis, from the Polarized Growth of Hyphae to Complex Reproduction and Infection StructuresMicrobiology and molecular biology reviews: MMBR 82:e00068–17https://doi.org/10.1128/MMBR.00068-17Google Scholar
- Glucose-induced cAMP signalling in yeast requires both a G-protein coupled receptor system for extracellular glucose detection and a separable hexose kinase-dependent sensing processMolecular microbiology 38:348–358https://doi.org/10.1046/j.1365-2958.2000.02125.xGoogle Scholar
- Feedback-regulated degradation of the transcriptional activator Met4 is triggered by the SCF(Met30)complexThe EMBO journal 19:282–294https://doi.org/10.1093/emboj/19.2.282Google Scholar
- Global gene deletion analysis exploring yeast filamentous growthScience 337:1353–1356https://doi.org/10.1126/science.1224339Google Scholar
- Gpa2, a G-protein alpha subunit required for hyphal development in Candida albicansEukaryotic cell 1:865–874https://doi.org/10.1128/EC.1.6.865-874.2002Google Scholar
- High efficiency transformation of intact yeast cells using single stranded nucleic acids as a carrierCurrent genetics 16:339–346https://doi.org/10.1007/BF00340712Google Scholar
- The unfolded protein response represses nitrogen-starvation induced developmental differentiation in yeastGenes & development 14:2962–2975https://doi.org/10.1101/gad.852300Google Scholar
- Experimental in Vivo Models of CandidiasisJournal of Fungi 4:21https://doi.org/10.3390/jof4010021Google Scholar
- Parallels in fungal pathogenesis on plant and animal hostsEukaryotic cell 5:1941–1949https://doi.org/10.1128/EC.00277-06Google Scholar
- HWP1 functions in the morphological development of Candida albicans downstream of EFG1, TUP1, and RBF1Journal of bacteriology 181:5273–5279https://doi.org/10.1128/JB.181.17.5273-5279.1999Google Scholar
- Moxifloxacin-Mediated Killing of Mycobacterium tuberculosis Involves Respiratory Downshift, Reductive Stress, and Accumulation of Reactive Oxygen SpeciesAntimicrobial agents and chemotherapy 66:e0059222https://doi.org/10.1128/aac.00592-22Google Scholar
- Modulation of the complex regulatory network for methionine biosynthesis in fungiGenetics 217:iyaa049https://doi.org/10.1093/genetics/iyaa049Google Scholar
- Microbial synergy via an ethanol-triggered pathwayMolecular and cellular biology 24:3874–3884https://doi.org/10.1128/MCB.24.9.3874-3884.2004Google Scholar
- The abundance of Met30p limits SCF(Met30p) complex activity and is regulated by methionine availabilityMolecular and cellular biology 20:7845–7852https://doi.org/10.1128/MCB.20.21.7845-7852.2000Google Scholar
- ADH1 promotes Candida albicans pathogenicity by stimulating oxidative phosphorylationInternational journal of medical microbiology: IJMM 309:151330https://doi.org/10.1016/j.ijmm.2019.151330Google Scholar
- Ime1 and Ime2 are required for pseudohyphal growth of Saccharomyces cerevisiae on nonfermentable carbon sourcesMolecular and cellular biology 30:5514–5530https://doi.org/10.1128/MCB.00390-10Google Scholar
- Hyphal induction under the condition without inoculation in Candida albicans is triggered by Brg1-mediated removal of NRG1 inhibitionMolecular microbiology 108:410–423https://doi.org/10.1111/mmi.13944Google Scholar
- A dominant suppressor mutation of the met30 cell cycle defect suggests regulation of the Saccharomyces cerevisiae Met4-Cbf1 transcription complex by Met32The Journal of biological chemistry 283:11615–11624https://doi.org/10.1074/jbc.M708230200Google Scholar
- Metabolism of sulfur amino acids in Saccharomyces cerevisiaeMicrobiology and molecular biology reviews: MMBR 61:503–532https://doi.org/10.1128/mmbr.61.4.503-532.1997Google Scholar
- Coevolution of morphology and virulence in Candida speciesEukaryotic cell 10:1173–1182https://doi.org/10.1128/EC.05085-11Google Scholar
- The pathogen Candida albicans hijacks pyroptosis for escape from macrophagesmBio 5:e00003–e14https://doi.org/10.1128/mBio.00003-14Google Scholar
- Cyclic AMP-protein kinase A and Snf1 signaling mechanisms underlie the superior potency of sucrose for induction of filamentation in Saccharomyces cerevisiaeEukaryotic cell 7:286–293https://doi.org/10.1128/EC.00276-07Google Scholar
- Morphogenesis in fungal pathogenicity: shape, size, and surfacePLoS pathogens 8:e1003027https://doi.org/10.1371/journal.ppat.1003027Google Scholar
- The Missing Link between Candida albicans Hyphal Morphogenesis and Host Cell DamagePLoS pathogens 12:e1005867https://doi.org/10.1371/journal.ppat.1005867Google Scholar
- Development and validation of real-time quantitative reverse transcriptase-polymerase chain reaction for monitoring gene expression in cardiac myocytes in vitroAnalytical biochemistry 270:41–49https://doi.org/10.1006/abio.1999.4085Google Scholar
- Global analysis of nutrient control of gene expression in Saccharomyces cerevisiae during growth and starvationProceedings of the National Academy of Sciences of the United States of America 101:3148–3153https://doi.org/10.1073/pnas.0308321100Google Scholar
- Glycolysis regulates gene expression by promoting the crosstalk between H3K4 trimethylation and H3K14 acetylation in Saccharomyces cerevisiaeJournal of genetics and genomics = Yi chuan xue bao 46:561–574https://doi.org/10.1016/j.jgg.2019.11.007Google Scholar
- GPR1 encodes a putative G protein-coupled receptor that associates with the Gpa2p Galpha subunit and functions in a Ras-independent pathwayThe EMBO journal 17:1996–2007https://doi.org/10.1093/emboj/17.7.1996Google Scholar
- The yeast ubiquitin ligase SCFMet30 regulates heavy metal responseMolecular biology of the cell 16:1872–1882https://doi.org/10.1091/mbc.e04-12-1130Google Scholar
- AMP deaminase as a control system of glycolysis in yeast. Mechanism of the inhibition of glycolysis by fatty acid and citrateThe Journal of biological chemistry 257:10644–10649Google Scholar
- G-protein coupled receptor from yeast Saccharomyces cerevisiaeBiochemical and biophysical research communications 240:287–292https://doi.org/10.1006/bbrc.1997.7649Google Scholar
Article and author information
Author information
Version history
- Preprint posted:
- Sent for peer review:
- Reviewed Preprint version 1:
Cite all versions
You can cite all versions using the DOI https://doi.org/10.7554/eLife.109075. This DOI represents all versions, and will always resolve to the latest one.
Copyright
© 2025, Shah et al.
This article is distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use and redistribution provided that the original author and source are credited.
Metrics
- views
- 0
- downloads
- 0
- citations
- 0
Views, downloads and citations are aggregated across all versions of this paper published by eLife.