Cell growth in the different conditions tested
S. cerevisiae strain BY4742 was grown on either YPD or SD medium with CdS QD concentrations of 25 to 200 mg L− 1. The colony spot assay showed that yeast cells grew better on YPD than on SD medium, therefore YPD was chosen for all subsequent experiments (Fig. 1A). When nystatin was added at 0.25 mg L− 1 (29) growth curve assays comparable to the control in YPD were obtained in the presence of 100 mg L− 1 of CdS QDs (Fig. 1B). The concentration of 100 mg L− 1 CdS QDs, with and without 0.25 mg L− 1 nystatin, was chosen as the treatment for subsequent analyses. (25, 29) The growth and treatment selected were identical to those used in previous transcriptomics analyses, (25, 28) allowing comparison between affected transcripts and proteins upon treatment with CdS QDs. Duration of the treatment was first set at 9 h, which corresponds to the exponential growth phase of the yeast cultures, and then at 24 h for the stationary phase. Cell cultures sampled at the exponential phase showed an OD600 value of about 2.5 for the control and 0.6 for QDs treatment, whereas cultures harvested at the stationary phase showed an OD600 value of about 12.0 for the control and 4.5 for QDs treatment with and without nystatin.
Proteomic variations in response to CdS QDs
Qualitative and quantitative changes in the yeast proteome during CdS QDs treatments were obtained from the 2D-PAGE-based and gel-free iTRAQ approaches, respectively. (45, 46) Common proteins identified with the two methods and after the two times of treatments are represented in Figs. 2 and 3 and in Table 1.
Table 1
List of the differentially expressed proteins after 9 h treatment 2D-gel (A), iTRAQ (B), and 24 h treatment (iTRAQ) (C).
| Proteins | Gene | pathway code | pathway name | description |
B C | 3-isopropylmalate dehydratase | LEU1 | sce00290 sce01100 sce01110 sce01230 | Valine leucine and isoleucine biosynthesis Metabolic pathway Biosynthesis of secondary metabolites Biosynthesis of amino acids | Catalyzes the isomerization between 2-isopropylmalate and 3-isopropylmalate, via the formation of 2-isopropylmaleate. |
B C | 78 kDa glucose-regulated protein homolog | KAR2 | sce03060 sce04141 | Protein export Protein processing in endoplasmic reticulum | Probably plays a role in facilitating the assembly of multimeric protein complexes inside the ER. Is required for secretory polypeptide translocation. |
B C | Adenylosuccinate lyase | ADE13 | sce00230 sce00250 sce01100 sce01110 sce01130 | Purine metabolism Alanine, aspartate and glutamate metabolism Metabolism secondary Biosynthesis of secondary metabolites Biosynthesis of antibiotics | This protein is involved in the subpathway that synthesizes AMP from IMP. |
A B C | ATP-dependent molecular chaperone HSC82 | HSC82 | sce04141 | Protein processing in endoplasmic reticulum | Molecular chaperone that promotes the maturation, structural maintenance and proper regulation of specific target proteins involved in cell cycle control and signal transduction. Undergoes a functional cycle that is linked to its ATPase activity. Interacts dynamically with various co-chaperones that modulate its substrate recognition, ATPase cycle and chaperone function. |
B C | Bifunctional purine biosynthesis protein ADE17 | ADE17 | sce00230 sce00670 sce01110 sce01130 | Purine metabolism One carbon pool by folate Secondary metabolites Biosynthesis of antibiotics | This protein is involved in the subpathway that synthesizes 5-formamido-1-(5-phospho-D-ribosyl)imidazole-4-carboxamide from 5-amino-1-(5-phospho-D-ribosyl)imidazole-4-carboxamide (10-formyl THF route). |
A C | Carnitine O-acetyltransferase, mitochondrial | CAT2 | sce04146 | Peroxisome | Carnitine acetylase is specific for short chain fatty acids. It seems to affect the flux through the pyruvate dehydrogenase complex. It may be involved as well in the transport of acetyl-CoA into mitochondria. |
B C | Cystathionine gamma-lyase | CYS3 | sce00260 sce00270 sce00450 sce01100 sce01130 sce01230 | Glycine sereine and threonine metabolism Cysteine and methionine metabolism Seleno compound metabolism Metabolic pathway Biosynthesis of antibiotics Biosynthesis of amino acid | This protein is involved in the subpathway that synthesizes L-cysteine from L-homocysteine and L-serine. |
B C | Elongation factor 1-beta | EFB1 | | | Catalytic subunit of the guanine nucleotide exchange factor (GEF) (eEF1B subcomplex) of the eukaryotic elongation factor 1 complex (eEF1). |
A C | Enolase 1 | ENO1 | sce00010 sce00680 sce01100 sce01110 sce01130 sce01200 sce01230 sce03018 | Glycolysis / Gluconeogenesis Methane metabolism Metabolic pathway Biosynthesis of secondary metabolites Biosynthesis of antibiotics Carbon metabolism Biosynthesis of amino acid RNA degradation | This protein is involved in the subpathway that synthesizes pyruvate from D-glyceraldehyde 3-phosphate. |
B C | FACT complex subunit POB3 | POB3 | | | Component of the FACT complex, a general chromatin factor that acts to reorganize nucleosomes. |
B C | Flavohemoprotein | YHB1 | | | Is involved in NO detoxification in an aerobic process. |
A B C | Fructose-bisphosphate aldolase | FBA1 | sce00010 sce00030 sce00051 sce00680 sce01100 sce01110 sce01130 sce01200 sce01230 | Glycolysis / Gluconeogenesis pentose phosphate pathway fructose and mannose metabolism methane metabolism metabolic pathway biosynthesis of secondary metabolites biosynthesis of antibiotics carbon metabolism biosynthesis of amino acid | Catalyzes the aldol condensation of dihydroxyacetone phosphate (DHAP or glycerone-phosphate) with glyceraldehyde 3-phosphate (G3P) to form fructose 1,6-bisphosphate (FBP) in gluconeogenesis and the reverse reaction in glycolysis. |
B C | Glutamate synthase [NADH] | GLT1 | sce00250 sce00910 sce01100 sce01110 sce01130 sce01230 | Alanine aspartate glutamate metabolism Nitrogen metabolism Metabolic pathway Secondary metabolites Biosynthesis of antibiotics Biosynthesis of amino acid | Forms L-glutamate from L-glutamine and 2-oxoglutarate |
B C | Glutamine synthetase | GLN1 | sce00250 sce00220 sce00630 sce00910 sce01100 sce01230 | Alanine aspartate glutamate metabolism Arginine biosynthesis Glyoxylate and dicarboxylate metabolism Nitrogen pathway Metabolic pathway Biosynthesis of amino acid | ATP binding and glutamate-ammonia ligase activity |
B C | Heat shock protein 26 | HSP26 | sce04141 | Protein processing in endoplasmic reticulum | One of the major polypeptides produced on heat shock. |
B C | Heat shock protein SSA1 | SSA1 | sce04140 sce04141 sce04144 sce04213 | Spliceosome Protein processing in endoplasmic reticulum Endocytosis Longevity regulating pathway | May play a role in the transport of polypeptides both across the mitochondrial membranes and into the endoplasmic reticulum. |
A B C | Homocysteine/cysteine synthase | MET17 | sce00270 sce00920 sce01100 sce01110 sce01130 sce01200 sce01230 | Cysteine and methionine metabolism Sulphur metabolism Metabolic pathway Biosynthesis of secondary metabolites Biosynthesis of antibiotics Carbon metabolism Biosynthesis of amino acid | Catalyzes the conversion of O-acetyl-L-homoserine (OAH) into homocysteine in the methionine biosynthesis pathway. |
B C | NADPH-dependent alpha-keto amide reductase | YDL124W | | | Involved in mitochondrial protection of cadmium-induced oxidative stress. |
B C | NAD-specific glutamate dehydrogenase | GDH2 | sce00220 sce00250 sce00910 sce01100 | Arginine biosynthesis Arginine alanine glutamine biosynthesis Nitrogen metabolism Metabolic pathway | NAD+-dependent glutamate dehydrogenase which degrades glutamate to ammonia and alpha-ketoglutarate. |
B C | Peptidyl-prolyl cis-trans isomerase D | CPR5 | | | PIases accelerate the folding of proteins. It catalyzes the cis-trans isomerization of proline imidic peptide bonds in oligopeptides. |
B C | Peroxiredoxin TSA1 | TSA1 | | | Thiol-specific peroxidase that catalyzes the reduction of hydrogen peroxide and organic hydroperoxides to water and alcohols, respectively. |
B C | Potassium-activated aldehyde dehydrogenase, mitochondrial | ALD4 | sce00010 sce00071 sce00280 sce00310 sce00620 sce01110 | Glycolysis / Gluconeogenesis Fatty acid degradation Valine leucine isoleucine degradation Lysine degradation Pyruvate metabolism Metabolic pathway | Potassium-activated aldehyde dehydrogenase involved in acetate formation during anaerobic growth on glucose. |
A C | Protein HBT1 | | | | Polarity-determining protein which forms a conjugate with the ubiquitin-like modifier HUB1. Involved in bud site selection and cellular morphogenesis during conjugation. |
B C | Ribosomal protein L15 | K7_RPL15A | | | structural constituent of ribosome |
B C | S-adenosylmethionine synthase 2 | SAM2 | sce00270 sce01100 sce01110 sce01230 | Cysteine and methionine metabolism Metabolic pathway Secondary metabolites Amino acid biosynthesis | Catalyzes the formation of S-adenosylmethionine from methionine and ATP. |
B C | Serine hydroxymethyltransferase, cytosolic | SHM2 | sce00260 sce00460 sce00630 sce00670 sce00680 sce01100 sce01110 sce01130 sce01200 sce01230 | Glycine serine and threonine metabolism Cyanoamino acid metabolism Glyoxylate and dicarboxylate metabolism One carbon pool by folate Methane metabolism Metabolic pathways Biosynthesis of secondary metabolites Biosynthesis of antibiotics Carbon metabolism Biosynthesis of amino acids | Interconversion of serine and glycine. |
B C | Succinate dehydrogenase [ubiquinone] flavoprotein subunit, mitochondrial | SDH1 | sce00020 sce00190 sce01100 sce01110 sce01130 sce01200 | Citrate cycle Oxidative phosphorylation Metabolic pathway Biosynthesis of secondary metabolites Biosynthesis of antibiotics Carbon metabolism | Catalytic subunit of succinate dehydrogenase (SDH) that is involved in complex II of the mitochondrial electron transport chain and is responsible for transferring electrons from succinate to ubiquinone (coenzyme Q). |
B C | Sulfite reductase [NADPH] subunit beta | MET5 | sce00920 | Sulfur metabolism | Catalyzes the reduction of sulfite to sulfide, one of several activities required for the biosynthesis of L-cysteine from sulfate. |
B C | UBX domain-containing protein 1 | SHP1 | sce04101 | Protein processing in endoplasmic reticulum | Involved in CDC48-dependent protein degradation through the ubiquitin/proteasome pathway. |
A B C | Uncharacterized oxidoreductase YMR226C | YMR226C | sce00240 sce00260 sce01100 | Pyrimidine metabolism Glycine serine threonine biosynthesis Metabolic pathway | NADP-dependent dehydrogenase with broad substrate specificity acting on 3-hydroxy acids. |
B C | Uracil phosphoribosyltransferase | FUR1 | sce00240 sce01100 | Pyrimidine metabolism, Metabolic pathway | Catalyzes the conversion of uracil and 5-phospho-alpha-D-ribose 1-diphosphate (PRPP) to UMP and diphosphate. |
Identification of differentially expressed proteins with 2D-PAGE
For the yeast cells exposed to CdS QDs for 9 h, with and without nystatin, the 2D-PAGE approach allowed the visualisation of around 900 spots for each sample. Subsequent MALDI-TOF MS-MS analysis allowed the identification of about 270 spots (Figure S3). Within the former group, 100 spots varied in intensity in response to the treatments: 81 of these differed from the control (ctr, not treated) vs QDs; 78 were different between ctr and nystatin + QDs; 11 differed between ctr and nystatin; and 72 differed between the nystatin and nystatin + QDs samples.
The CdS QDs treatment, with and without nystatin, altered the expression level of 56 common proteins as found by comparing ctr and QDs; ctr and nystatin + QDs; and nystatin vs nystatin + QDs. Among the 11 proteins affected by the treatment with nystatin, only 4 were common to other treatments, and were thus identified, whilst those proteins that were present exclusively in the samples treated with nystatin were excluded from the analysis of the effect of CdS QDs (Fig. 2A). At 9 h, there is a balance in the number of the modulated proteins between up and downregulated (fig S7A). The identities of the protein spots, whose abundance was differentially modulated with a p value of ≤ 0.05 is presented in Table S1.
Identification of the differentially expressed proteins using iTRAQ labelling
The time points for quantitative iTRAQ analysis were 9 and 24 h. This gel-free approach allowed processing more samples than 2D-PAGE, therefore proteome variations were also analysed under all treatments and both time points. The iTRAQ approach enables quantification at the peptide level and direct protein mapping because both types of information originate from the same MS-MS spectra. In several other iTRAQ studies, about a thousand proteins were identified. (47 48) Far more than a thousand proteins were detected here within each single iTRAQ experiment on each biological replicate (Figure S4 and S5).
The iTRAQ experiments corresponding to 9 h of treatment allowed the identification and quantification of 1129 (934 quantified), and 1055 (835) unique proteins from the two biological replicates BR1 and BR2, respectively (Figure S4 and Supplementary Table S2). Of these, 849 (712) proteins were common to both biological replicates.
The iTRAQ analysis revealed 92 proteins enriched in the yeast cells in response to the treatments with CdS QDs, with and without nystatin: 62 of these proteins were identified by comparing ctr vs QDs; 45 were identified from the comparison of ctr and nystatin + QDs, 16 by comparing ctr vs nystatin and 59 by comparing the nystatin and nystatin + QDs samples. The CdS QD treatment altered the abundance level of 71 common proteins between ctr vs QDs, ctr vs nystatin + QDs and nystatin vs nystatin + QDs. Only 4 proteins from the comparison of ctr vs nystatin were identified because the other 10 proteins were not in common with any other treatment and therefore were not considered relevant. (Fig. 2B). Nevertheless, at 9 h the iTRAQ results complement the 2D-PAGE results by reconfirming the general trends and supporting the identification of specific protein clusters affected by CdS QD exposure (Figure S7B). The complementary nature of these methods was highly useful: proteins identified with 2D-PAGE and iTRAQ differ substantially as shown in Figure S6 and in Table 4, but the combined use of different techniques uncovers a higher proportion of the proteome of an organism. (49) The list of proteins identified in the two biological replicates and the proteins common to all samples are shown in Tables S2 and S4.
The iTRAQ analysis of the 24 h samples allowed the identification of 943 (886 quantified), and 1346 (1080) unique proteins from the two biological replicates BR1 and BR2, respectively (Figure S5 and Supplementary Table S3). Of these, 562 (505) proteins were common to the two biological replicates.
The iTRAQ-based quantitative analysis revealed that the total number of proteins enriched in the yeast cells in response to all treatments with CdS QDs, with and without nystatin, was 127. Eighty- eight of these proteins were identified by comparing ctr vs QDs, 86 by comparing ctr and nystatin + QDs, 23 by comparing ctr vs nystatin and 94 from the difference between the nystatin and nystatin + QDs samples. The CdS QD treatment altered the expression level of 59 common proteins as judged from the comparison of ctr vs QDs, ctr vs nystatin + QDs, and nystatin vs nystatin + QDs. Only 15 proteins were identified from the comparison between ctr vs nystatin because the other 8 proteins were not common to the other treatments and were therefore not considered relevant (Fig. 2C and S7C). The lists of proteins identified in two biological replicates and proteins common to all datasets are shown in Tables S3 and S4. Notably, at 24 h the modulated proteins showed a different trend to the finding after 9 h of treatment, i.e. the majority of the proteins were downregulated (Figure S7C).
We pooled together all the proteins identified for the 9 h treatments obtained with both methods, and compared them with those obtained for the 24 h treatment resulting from the iTRAQ method. Four proteins were in common between the two methods at both 9 h and 24 h: ATP-dependent molecular chaperone Hsp82, uncharacterised oxidoreductase YMR226C, fructose-bisphosphate aldolase (Fba1), and homocysteine/cysteine synthase (Met17) (Fig. 3 and Table S4). However, another 4 proteins were in common between the 2D -gel method at 9 h and iTRAQ at 24 h. These proteins were: carnitine O-acetyltransferase mitochondrial (Cat2), folic acid synthesis protein (Fol1), serine/threonine-protein kinase Ypk1, and succinate dehydrogenase [ubiquinone] iron-sulphur subunit (Sdh2). Another 26 proteins were in common between the iTRAQ method at 9 h and iTRAQ at 24 h. The most downregulated proteins were elongation factor 1-beta (Efb1) and glutamate synthase (Glt1), and the most upregulated proteins were S-adenosylmethionine synthase 2 (Sam2) and superoxide dismutase 1 copper chaperone (Ccs1) (Fig. 3, Table 1 and Table S4).
Ontology analysis of the identified proteins
Analysis using gene ontology (GO) groups proteins based on biological processes, molecular functions, and cellular components. A GO term difference with a p < 0.05 was considered as significant enrichment. This annotation of proteins into different classes was instrumental to understanding their biological relevance. Functional in silico classification of the 192 proteins obtained at 9 h with both 2D-PAGE and iTRAQ labelling and the 119 proteins identified with iTRAQ at 24 h, was achieved via GO analysis using the software PANTHER. A total of 39 slim GO terms were significantly enriched (p < 0.05). PANTHER grouped all the enriched proteins at 9 h into four groups based on their molecular functions (Figure S8A): small molecule binding (4.88%), oxidoreductase activity (3.95%), structural constituent of ribosome (3.17%) and catalytic activity (1.75%). The major molecular GO functions for the 24 h samples were: oxidoreductase activity (6.03%), proton transmembrane transporter activity (5.02%), and catalytic activity (2.10%) (Figure S8B).
When the enriched proteins identified at 9 h were analysed on the basis of biological processes, they were organized in five major groups: organic acid biosynthetic process (4.82%), carboxylic acid biosynthetic process (4.82%), carbohydrate metabolic process (4.69%), primary metabolic process (3.47%), and metabolic process (1.70%). The enriched proteins obtained for the 24 h treatment were subdivided in 15 groups, of which the more important were: aerobic respiration (15.63%), tricarboxylic acid cycle (15.05%), cellular respiration (13.04%), energy derivation by oxidation of organic compounds (13.04%), oxidation-reduction process (10.67%), ATP synthesis coupled proton transport (9.23%), energy coupled proton transport (9.23%), and carbohydrate metabolic process (5.28%) (Fig. 4A and B). The main GO cell component categories for the 9 h treatment were: cytosolic ribosome (3.45%) and cytosolic part (3.15%). The samples recovered after 24 h treatment were enriched in mitochondrial inner membrane (5.08%) and cytoplasmic (1.61%) proteins (Figure S8C and S8D).
GO analysis of the differentially abundant proteins identified ‘oxidoreductase activity’ and ‘catalytic activity’ as the most perturbed biochemical functions in response to CdS QD exposure at 9 and 24 h, whilst the GO biological processes that differ between the two times of exposure corresponded to ‘carbohydrate metabolic process’ and ‘metabolic process’. Analysis of the significant biological processes affected at 24 h revealed that the majority of the GO classes were downregulated, in particular aerobic respiration, tricarboxylic acid cycle, cellular respiration, energy derivation by oxidation of organic compounds, oxidation-reduction process, and ATP synthesis coupled proton transport. Overall these results show that the treatment with CdS QDs is time-dependent.
In particular, two of the downregulated proteins that belong to each of the aerobic respiration, cellular respiration and tricarboxylic acid (TCA) cycle classes were citrate synthases CIT1 and CIT2. In eukaryotes, the TCA cycle occurs in the mitochondrial matrix and plays a pivotal role in the utilization of non-fermentable carbon sources via oxidative generation of reducing equivalents (NADH), driving aerobic respiration to yield ATP. (50) The TCA cycle is also an important source of biosynthetic building blocks, such as α-ketoglutarate, succinyl-CoA and oxaloacetate, which are required for the synthesis of glucose and amino acids. (50) Yeasts have multiple citrate synthase genes (CIT1, CIT2, and CIT3), but it is not clear how they differ in function or if any of them encode a specific methylcitrate synthase. The products of the CIT1 and CIT3 genes have been shown to be mitochondrial proteins, whereas that of the CIT2 gene is clearly peroxisomal. (51)
The foregoing molecular function and biological processes mostly linked to mitochondrial function and structure represent the “core response” to CdS QDs. These data are in keeping with other results obtained from simple eukaryotic organisms and human cell lines. (25, 28, 26) From a physiological and molecular point of view, it has been demonstrated that ENMs increase ROS production by interacting negatively with all cell compartments, in particular by affecting cell membranes and the mitochondria and, consequently, the levels of energy production and cellular respiration. (25) The correspondence between ROS production and inhibition of respiration has been reported in the literature. For example, Fe3O4 nanoparticles have an inhibitory effect on yeast growth. The inhibition is attributed to their interaction with the mitochondria, leading to disruption of the mitochondrial respiratory chain complex IV, and consequent attenuation of ATP production. (52) In addition, it has been found that NiO NPs inhibit metabolic activity, induce intracellular accumulation of ROS, and provoke cell death in S. cerevisiae. (53)
Pathway analysis of the identified proteins
Metabolic pathway analysis was performed by submitting the Gene IDs of the proteins to the KEGG server (http://www.kegg.jp) for S. cerevisiae to identify the pathways that were represented more frequently. At 9 h the main pathway classes were: general metabolic pathway, biosynthesis of secondary metabolites, biosynthesis of amino acids, glycolysis and gluconeogenesis, protein biosynthesis, carbon metabolism, and protein processing in endoplasmic reticulum (ER) (Fig. 5).
At 24 h the main pathway classes were: general metabolic pathway, biosynthesis of secondary metabolites, oxidative phosphorylation, TCA cycle, glycolysis and gluconeogenesis, pyruvate metabolism, protein biosynthesis, carbon metabolism, and protein processing in endoplasmic reticulum (ER) (Fig. 5).
Of particular interest was the pathway “glycolysis and gluconeogenesis”, common to the two treatment times (Fig. 6), which included 13 proteins identified at 9 h, and 11 at 24 h. At 9 h of treatment, eight enzymes associated to the glycolysis pathway were over-abundant: NADP-dependent alcohol dehydrogenase 6 (Adh6), NADP-dependent alcohol dehydrogenase 7 (Adh7), glyceraldehyde-3-phosphate dehydrogenase 1 (Tdh1), glyceraldehyde-3-phosphate dehydrogenase 3 (Tdh3), glucose-6-phosphate isomerase (Pgi1), glucokinase-1 (Glk1), mitochondrial pyruvate dehydrogenase complex protein X component (Pdx1), and mitochondrial potassium-activated aldehyde dehydrogenase (Ald4). Five enzymes were under-abundant: fructose-bisphosphate aldolase (Fba1), enolase 1 (Eno1), enolase 2 (Eno2), pyruvate decarboxylase isozyme 1 (Pdc1), pyruvate decarboxylase isozyme 1(Pfk2).
At 24 h of treatment, the majority of the enzymes associated the glycolytic pathway were under-abundant: fructose-1,6-bisphosphatase (Fbp1), fructose-bisphosphate aldolase (Fba1), enolase 1 (Eno1), acetyl-coenzyme A synthetase 1 (Acs1), dihydrolipoyl dehydrogenase (Ldp1), mitochondrial dihydrolipoyl dehydrogenase (Ldp1), pyruvate decarboxylase isozyme 5 (Pdc5), mitochondrial potassium-activated aldehyde dehydrogenase (Ald4), and NADP-dependent alcohol dehydrogenase 2 (Adh2). Only two enzymes were detected at levels higher than the control: hexokinase-1 (Hxk1) and glyceraldehyde-3-phosphate dehydrogenase 2 (Tdh2).
Among the enzymes common between 9 h and 24 h, Fba1 and Eno1 were under-abundant at both times of treatment, whereas the levels of Ald4 were initially increased at 9 h, then decreased at 24 h. As reported by Gomes et al. (2006) ENMs treatment inhibited the glycolytic pathway and stimulated fermentation. (54) Horstmann et al., (2019) suggested as highly probable that sugar transport genes and sugar-utilising enzyme genes are simultaneously affected by the presence of Cd-QDs. (55) They conjecture that the ENO1 gene is downregulated as a consequence of transport of low levels of sugars caused by the suboptimal activity of glucose transporters due to the presence of Cd-QDs. Conversely, the three isoforms of glyceraldehyde-3-phosphate dehydrogenase, GAPDH (Tdh1, Tdh2, Tdh3), were found to be upregulated for both treatment times. GAPDH is a glycolytic enzyme involved in glucose degradation and energy yield. It catalyses the sixth step of glycolysis, i.e. the conversion of glyceraldehyde-3-phosphate to 1,3 bis-phosphoglycerate, but also displays non-glycolytic activity in certain subcellular locations. In vitro inhibition studies of GAPDH in the presence of QDs suggest that binding of QDs to the enzyme molecules slows down the rate of the reactions catalysed by the enzyme, suggesting that QDs may act as enzyme inhibitors. (56) When human cancer cells are exposed to QDs, the loss of cellular GAPDH activity causes a metabolic perturbation during glycolysis, and the inhibition of GAPDH leads to the decrease of glycolytic rates. This suggests a possible mechanism of change in energy production from the glycolytic pathway to fermentation during QD-mediated cellular injury. This process may lead eventually to cell dysfunction and death. (56)
Proteins leading to the Krebs cycle (Pdx1, Acs1, Lpd1, Ald4) or to fermentation (Adh2, Adh6, Adh7, Pdc1, Pdc5) were modulated during treatment with CdS QDs for both periods (Table S4 and Fig. 6). Pdc1 is the most prevalent form of the three yeast pyruvate decarboxylases which are involved both in the anaerobic fermentation of pyruvate to acetaldehyde and in amino acid catabolism. Pdc1, together with Tdh2 and Tdh3, was found among the proteins that constitute the hard corona in yeast during CdS QDs treatments, with a specific role in determining the toxicity of these ENMs. (57)
Another pathway of particular interest is “protein processing in ER”, which includes 7 proteins modulated at 9 h (2 under-abundant and 5 over-abundant) and 6 at 24 h (2 proteins with reduced levels and 3 with increased levels) (Fig. 7). Three common enzymes were found to be overexpressed at the two times of treatments: ATP-dependent molecular chaperone Hsc82, 78 kDa glucose-regulated protein homolog (Kar2) and UBX domain-containing protein 1 (Shp1). Hsc82, a member of the Hsp90 family, acts to promote the maturation, structural maintenance and regulation of proteins involved in cell cycle control, ribosome stability and signal transduction. (58) Hsp90 proteins operate in a number of signalling pathways which are altered during exposure to metal ENMs. (59) It was shown that Hsc82 is one of the main hubs in CdS QDs sensitivity, (28) and that it is one of the hard corona proteins for CdS QDs. (57)
Other two enzymes of the ER were present at higher levels at 9 h and lower levels at 24 h. These enzymes were heat shock protein 26 (Hsp 26) and heat shock protein Ssa1, which is a ribosome-associated member of the Hsp70 family participating in the folding of newly-synthesized polypeptides. (60) In addition, the cell division control protein 48 (Cdc48) was less abundant at 9 h.
The results obtained by Wei et al. (2017) on human cancer cells suggest that some ENMs are capable of inducing autophagy and affecting the ER. (61) Schütz et al. (2016) reported that internalized silica nanoparticles (Si-NPs) may accumulate in lysosomes, resulting in lysosomal dysfunction in HeLa cells. (62) Similarly, Si-NPs accumulating in ER indicate an effect on ER structure, through mechanisms still unknown. Any damage to ER is closely connected with cell autophagy, one of the principal cell death mechanisms triggered by ENMs. The acute toxicity of ZnO NPs to Daphnia pulex evidenced by proteomic results showed that some processes such as protein synthesis and translocation across the ER were inhibited to reduce the stress associated to protein misfolding. (63) The majority of the modulated proteins involved in “oxidative phosphorylation” are from the 24 h treatment with CdS QDs, except for the mitochondrial succinate dehydrogenase (ubiquinone) flavoprotein subunit which responded to treatment with CdS QDs at 9 h. Ten of the proteins found in changed amounts after the 24 h treatment with CdS QDs were downregulated suggesting energy production was significantly lessened. These proteins are the mitochondrial succinate dehydrogenase (ubiquinone) iron-sulphur subunit (Sdh2), cytochrome b-c1 complex subunit 6 (Qcr6), the mitochondrial cytochrome b-c1 complex subunit 7 (Qcr7), cytochrome b-c1 complex subunit (Rip1, the Rieske protein), cytochrome c oxidase subunit 4 (Cox4), the mitochondrial cytochrome c oxidase subunit 6 (Cox6), the mitochondrial ATP synthase subunits 5, d, gamma, and delta (Atp5, Atp16, Atp3 and Atp16), while cytochrome c oxidase subunit 2 (Cox2) and the vacuolar isoform of the V-type proton ATPase subunit a (Vph1) were both upregulated. It appears that after 24 h of treatment most of the mitochondrial proteins had reduced activity, causing a slow-down in oxidative phosphorylation and ATP production (Fig. 8). The proteins most affected by the CdS-QDs are components of mitochondrial respiration complexes III, IV and V. Mitochondria are a significant organelle in QD-induced toxicity. (64, 65) It has been shown that CdS QDs damage mitochondrial functionality and reduce respiration activity in yeast, (29) plants, (25) and human cells. (26) Damage to mitochondrial functions and structure caused by several types of metal-ENMs has been found in mollusc bivalve and mouse cells. (66, 67) Interestingly all the proteins of the ATP synthase complex were downregulated, which indicates a reduction in the energy produced through oxidative phosphorylation, and connects with a general downregulation of the enzymes involved in the glycolytic pathway.
In summary, the upregulation of fermentation, but downregulation in the levels of glucose, manifests as a change into lactate or acetate to provide enough energy for survival. These results might bypass the imbalance in the aerobic metabolism and the TCA cycle. Moreover, acetate is also regarded as an expedient source of energy for stressed cells. (68) These observations are consistent with the reports in which silver nanoparticles caused oxidative stress and defects in mitochondrial and endoplasmic reticulum (ER) enzymes. (69, 70) In aerobic metabolism, the formation of ROS is a natural by-product, but an excess of ROS can chemically modify proteins and lipids by lipid peroxidation and oxidative stress, thus leading to damage to vital organelles such as mitochondria, the ER, and lysosomes. (71, 72)
Inhibition of GAPDH activity by CdS QDs
Figure 9 shows that at both 9 and 24 h the activity of GAPDH in yeast cells treated with 100 mg L− 1 of CdS QDs was significantly lower than in the untreated samples (Fig. 9). Though not highly significant, the activity of GAPDH at 9 h was higher than at 24 h. Overall the CdS QDs treatment at both time points inhibits the glycolytic process at the level of the enzyme GAPDH, as suggested by the proteomic approach (Fig. 6). CdS QD treatment consistently altered GAPDH abundance and decreased GAPDH activity. In vitro experiments in the BY4742 yeast strain on hard corona demonstrated a strong dose-dependent reduction of the enzyme activity upon CdS QDs treatment. (57) The reduction of GAPDH activity by CdS QDs could be explained by CdS QD oxidation of the GAPDH active site (cysteine 152), which is known to lower GAPDH activity and reduce the accessibility to substrates such as glyceraldehyde-3-phosphate. (56, 57) ENPs can induce unfolding and a reduced activity of the identified proteins, as observed in the case of GAPDH isoforms, but CdS QD binding to hard corona proteins could mediate non-specific interactions with other cellular components. (56, 57)
Effect of CdS QDs on ROS generation and cell integrity in S. cerevisiae
Flow cytometry analysis showed that exposure for 9 h to CdS QDs led to a very substantial overproduction of ROS, while significant but much lower ROS overproduction was observed after 24 h of treatment (CdS QDs 100 mg L− 1). The results indicate that growth inhibition induced by the treatment was associated with oxidative stress having intense cytotoxic effects at 9 h. Figure 10A and Figure S9 show the time-dependent changes in intracellular production of ROS compared to the untreated control.
Production of ROS by nanomaterials is considered a major factor in QDs toxicity. The deleterious action of oxidative stress starts by causing oxidative damage to biomolecules and destroying their structure, which decreases cellular defences and ultimately leads to cell death, possibly by a mechanism more similar to apoptosis. (73) Overall, our data demonstrate that QDs change the expression levels of a number of proteins by inducing oxidative stress at both treatment periods. Therefore, it is possible to correlate the dysfunction in the glycolysis pathway, the downregulation of oxidative phosphorylation and also the increase in protein misfolding in the ER, all caused by QD treatment, with the production of ROS, which impairs the oxidative balance of the cells and becomes increasingly severe with time. (74, 75)
Figure 10B shows that after 9 h of CdS QDs treatment, the proportion of dead cells was 30% higher with respect to the control, whilst at 24 h the proportion of dead cells increased to 54%. Together these results indicate that cell death increased with the time and dose of CdS QDs. (76, 75)
Robustness of markers identification using multiomic approaches
The proteins that were up or down regulated following CdS QDs treatment were assessed against other omics markers identified using transcriptomics and phenomics, as reported elsewhere analyses. (28, 29, 25) Fig. 11 shows the levels of correlation between proteomics/transcriptomics, phenomics/transcriptomics and proteomics/phenomics markers. These data were obtained by comparing 284 significant proteins against more than 5000 haploid deletion mutants and the whole set of transcripts obtained with a yeast microarray platform. (28) The correspondences, both symmetric (++/--) and antisymmetric (+-/-+), consisted of a small percentage of the compared elements, i.e. 22 proteins, 14 transcripts, and 8 mutants which responded as were up/ down regulated and/or sensitive/tolerant to the treatment with CdS QDs. It is well known that the correspondence between proteomics and transcriptomics is typically low. (77) The molecular markers that showed this level of correlation in the three comparisons are considered robust enough to be candidates for omics exposure markers. The functions which are most implicated are glycolysis cycle and protein processing in the endoplasmic reticulum. Across the proteins, transcripts and growth phenotypes, the only common element is FKS1, which encodes the catalytic subunit of the yeast 1,3-β-d-glucan synthase, relevant in the building of yeast cell wall and consequently in cell response to materials exposure.