Skip to main content

Identification of an essential regulator controlling the production of raw-starch-digesting glucoamylase in Penicillium oxalicum

Abstract

Background

Raw-starch-digesting glucoamylases (RSDGs) from filamentous fungi have great commercial values in starch processing; however, the regulatory mechanisms associated with their production in filamentous fungi remain unknown. Penicillium oxalicum HP7-1 isolated by our laboratory secretes RSDG with suitable properties but at low production levels. Here, we screened and identified novel regulators of RSDG gene expression in P. oxalicum through transcriptional profiling and genetic analyses.

Results

Penicillium oxalicum HP7-1 transcriptomes in the presence of glucose and starch, respectively, used as the sole carbon source were comparatively analyzed, resulting in screening of 23 candidate genes regulating the expression of RSDG genes. Following deletion of 15 of the candidate genes in the parental P. oxalicum strain ∆PoxKu70, enzymatic assays revealed five mutants exhibiting significant reduction in the production of raw-starch-digesting enzymes (RSDEs). The deleted genes (POX01907, POX03446, POX06509, POX07078, and POX09752), were the first report to regulate RSDE production of P. oxalicum. Further analysis revealed that ∆POX01907 lost the most RSDE production (83.4%), and that POX01907 regulated the expression of major amylase genes, including the RSDG gene POX01356/PoxGA15A, a glucoamylase gene POX02412, and the α-amylase gene POX09352/Amy13A, during the late-stage growth of P. oxalicum.

Conclusion

Our results revealed a novel essential regulatory gene POX01907 encoding a transcription factor in controlling the production of RSDE, regulating the expression of an important RSDG gene POX01356/PoxGA15A, in P. oxalicum. These results provide insight into the regulatory mechanism of fungal amylolytic enzyme production.

Background

Starch is an important renewable carbohydrate biosynthesized in large quantities through plant photosynthesis. Starch biorefinery can provide a variety of useful chemicals, including biofuels. In traditional starch biorefinery, amylases, including α-amylase (EC 3.2.1.1) and glucoamylase (EC 3.2.1.3), are used to hydrolyze starch to glucose [1]. To reduce the energy costs associated with traditional starch biorefinery, raw-starch-digesting enzymes (RSDEs), specifically raw-starch-digesting glucoamylases (RSDGs), represent promising alternatives capable of directly degrading raw starch granules into oligosaccharides or glucose below the gelatinization temperature of starch [2].

Native RSDEs are primarily produced by filamentous fungi, such as Aspergillus sp. [3, 4], Penicillium sp. [5], and Aureobasidium pullulans [6]. Recently, a novel RSDG (PoxGA15A) exhibiting suitable properties was identified in Penicillium oxalicum and showed remarkably broad pH stability and substrate specificity. Simultaneous saccharification and fermentation of either raw cassava or corn flour using the recombinant protein rPoxGA15A from Pichia pastoris combined with the presence of commercial α-amylase resulted in high fermentation efficiency (> 90%) [5]. However, both native PoxGA15A and rPoxGA15A production, as well as that of other RSDGs from fungi, such as Rhizopus sp. A-11 [7], Aspergillus fumigatus CFU-01 [8] and Laceyella sacchari LP175 [9], are too low, which limit their industrial application. Notably, the expression of fungal RSDG genes is strictly controlled by transcription factors (TFs) at the transcriptional level. Genetic engineering of fungal strains based on constructed TF-specific regulatory networks and targets represents an efficient method to improve RSDG production.

Few studies associated with the regulation of RSDG gene expression in filamentous fungi, including Aspergillus sp., Talaromyces pinophilus, P. oxalicum, and Neurospora crassa have been undertaken. Previous studies described two Zn(II)2Cys6 zinc finger proteins (AmyR [10] and COL-26 [11]), a heterotrimeric G protein subunit (PGA3 [12]), and casein kinase CK2 proteins (CK2B1 and CK2B2) as activators of Amy15A (a PoxGA15A homolog) gene expression [13]. In addition to these activators, an extracellular protease activator, PrtT [14], and an HMG-box protein, PoxHmbB [15], were identified as Amy15A repressors in P. oxalicum. However, these findings are insufficient to elucidate the regulatory mechanism associated with RSDG gene expression for improving RSDG production.

In this study, we employed RNA-seq and molecular genetic technologies to screen and identify novel regulators of RSDE production and RSDG gene expression in P. oxalicum. Transcriptomes from P. oxalicum grown in the presence of glucose or soluble corn starch (SCS) were profiled to identify candidate regulators of RSDE production. Subsequent knockout of candidate genes, measurement of enzyme activity in the resulting mutants, and expression analysis of amylase genes, including the RSDG gene PoxGA15A, were performed to identify novel regulators of RSDG gene expression in P. oxalicum.

Results

Transcriptome profiling and screening of candidate regulators of RSDE production in P. oxalicum

Genome-wide screening of candidate regulators of RSDE production was undertaken through RNA-seq analysis of the transcriptome profiles of P. oxalicum grown on media containing glucose or SCS as the sole carbon source following a transfer from glucose. In the presence of glucose, carbon catabolite repression is activated and inhibits RSDE production in P. oxalicum, whereas SCS stimulates the secretion of RSDEs. Total RNAs were extracted from the mycelia of P. oxalicum grown on glucose or SCS for 4 h and then sequenced. Approximately 24 million clean reads at 100 bp in length (Accession Number SRP116594 in Sequence Read Archive [SRA] database) were generated from each sample, with > 90% mapped into the genome of P. oxalicum wild-type strain HP7-1 [16] (Additional file 1: Table S1). High Pearson’s correlation coefficients among three biological replicates of P. oxalicum under each culture condition (r ≥ 0.96) (Additional file 2: Figure S1) indicated the reliability of the transcriptome data. Gene-expression levels were quantitatively analyzed according to fragments per kilobase of exon per million mapped reads (FPKM), and the differences were evaluated using NOISeq [17].

Comparative transcriptome analyses identified 961 differentially expressed genes (DEGs), with 681 upregulated (0.52 < log2 fold change < 10.18) and 280 downregulated (− 10.26 < log2 fold change < − 0.54) in P. oxalicum HP7-1 on SCS as compared with that in HP7-1 on glucose (Additional file 3: Table S2). Kyoto Encyclopedia of Genes and Genomes (KEGG) annotation revealed that these DEGs were mainly involved in metabolic pathways, including carbohydrate, amino acid, and energy metabolism, and translation (Fig. 1a).

Fig. 1
figure 1

Comparative analysis of transcriptomes from Penicillium oxalicum HP7-1 grown in the presence of glucose and soluble corn starch. a Kyoto encyclopedia of genes and genomes annotation of proteins encoded by differentially expressed genes (DEGs) between HP7-1 grown in soluble corn starch medium relative to that grown in glucose medium. DEGs were screened using a probability threshold of ≥ 0.6 and a |log2 fold change| of > 0.5. b Heatmap showing the transcription levels of DEGs encoding carbohydrate-active enzymes (CAZymes). c The transcription levels of DEGs encoding putative transcription factors

Of the DEGs, 48 were annotated to encode carbohydrate-active enzymes (CAZymes), including six auxiliary activity families, 18 glycoside hydrolase families, seven glycosyl transferase families, three carbohydrate esterase families, and four carbohydrate-binding-module families. Among these 48 DEGs, 28 were upregulated, with a log2 fold change from 0.72 to 5.03 in HP7-1 on SCS relative to that in HP7-1 on glucose. As expected, these included 60% of the amylase genes in the whole genome of P. oxalicum HP7-1, including a key RSDG gene (POX01356/PoxGA15A) and a glucoamylase gene (POX02412), a key α-amylase gene (POX09352/Amy13A), and three α-glucosidase genes (POX03741, POX03889, and POX06242). Surprisingly, the important genes involved in the degradation of chitin and plant-cell walls were also found in the upregulated gene set, including three chitinase genes (POX00089/ChiB1, POX03021, and POX07145), a chitooligosaccharide deacetylase gene (POX06577), a lytic polysaccharide monooxygenase gene (POX05968), and an endo-β-1,4-galactanase gene (POX06689). In contrast, among the 20 downregulated DEGs, five encoded plant-cell-wall-degrading enzymes (CWDEs), including an expansin-like protein-encoding gene POX01524, an endo-β-1,4-glucanase gene POX04137, two endo-β-1,4-xylanase genes (POX06783/Xyn11A and POX08990), and a xylosidase gene (POX05540), and a chitinase gene (POX07971), with all exhibiting log2 fold changes from − 1.52 to − 0.77. However, no amylase genes were included (Fig. 1b).

Additionally, comparative analyses also revealed 23 DEGs encoding putative TFs as candidate regulators (Additional file 4: Table S3), most of which contained at least one zinc finger domain (C2H2, GATA, Zn2Cys6, or DHHC). Eight candidate DEGs increased target transcript level by 72.2%–316.7% in HP7-1 on SCS relative to that in HP7-1 on glucose, whereas 15 genes showed lower transcript levels (by 36.5–64.7%) (Fig. 1c). POX04510, a protein homolog of AreA in Aspergillus nidulans, negatively controls cellulase production [18]. Moreover, POX02944 and POX03789 encode protein homologs of OefC and StuA, respectively, that regulate sporulation in filamentous fungi, including A. nidulans [19, 20]. Recently, POX04860 and POX05726 were found to regulate cellulase production in P. oxalicum [21].

Five novel regulators are required for RSDE production in P. oxalicum

To investigate the regulatory roles of these 23 candidate TF-encoding DEGs in RSDE production in P. oxalicum, homologous recombinant technology was employed for their deletion from the parental strain ∆PoxKu70 [16]. Eleven deletion mutants (∆POX00852, ∆POX01907, ∆POX03446, ∆POX03789, ∆POX05041, ∆POX06509, ∆POX07078, ∆POX07522, ∆POX07938, ∆POX09088, and ∆POX09752) were successfully constructed in this present study, and four deletion mutants (∆POX02944, ∆POX04860, ∆POX05726, and ∆POX06425) had been constructed in the previously published work [21]. These 11 newly constructed deletion mutants in this study were verified by polymerase chain reaction (PCR) using specific primers (Additional file 5: Figure S2a and Additional file 6: Table S3). Both ∆POX03789 and ∆POX02944 were unable to produce spores (data not shown), which was consistent with previously described results [19, 20]. Assays of RSDE activity undertaken on all of the deletion mutants, except for ∆POX03789 and ∆POX02944, grown on medium containing SCS as the sole carbon source for 4–6 days after direct inoculation revealed that five deletion mutants (∆POX01907, ∆POX03446, ∆POX06509, ∆POX07078, and ∆POX09752) showed significant reduction in RSDE production relative to the parental strain ∆PoxKu70; ranking from 30 to 83.4% (P ≤ 0.05, Student’s t test) (Fig. 2a). This represents the first report showing that these five DEGs were involved in RSDE production in P. oxalicum (Table 1). Strikingly, the mutant ∆POX01907 showed higher losses of RSDE activity (83.4% at day 4 and 80.0% at day 6) relative to the other four mutants and was subsequently selected for further study. To exclude the possibility that multiple copies of POX01907-deletion cassette were integrated into the ∆PoxKu70 genome, the mutant ∆POX01907 was further confirmed by Southern hybridization analysis (Additional file 5: Figure S2b) using specific probes (Additional file 6: Table S3).

Fig. 2
figure 2

Screening and identification of novel regulatory genes required for the production of raw starch-degrading enzymes (RSDEs) in Penicillium oxalicum. a RSDE activities in strains with deletion of candidate regulatory genes and grown in medium containing soluble corn starch as the sole carbon source for 4–6 days after direct inoculation. The parental strain ∆PoxKu70 was used a control. Asterisks indicate significant differences (**P < 0.01; *P < 0.05) between the deletion mutants and the parental strain ∆PoxKu70 according to Student’s t test. b, c RSDE activities in the deletion mutants ∆POX01907 and ∆PoxKu70 and the complementary strain CPOX01907. P. oxalicum strains were cultivated in medium containing soluble corn starch for 2–4 days after transfer from glucose. Asterisks indicate significant differences (**P < 0.01) between ∆POX01907 and ∆PoxKu70 or CPOX01907, as assessed by Student’s t test

Table 1 Novel genes encoding TFs involved in regulating the production of raw-starch-degrading enzymes in P. oxalicum

POX01907 regulates RSDE production in P. oxalicum following SCS induction

To elucidate POX01907-specific regulatory roles in RSDE production, P. oxalicum strains ∆POX01907 and ∆PoxKu70 were grown on medium containing SCS as the sole carbon source for 2–4 days after a transfer from glucose, followed by real-time investigation. The results revealed 43.5%–71.3% reduction in RSDE production by the mutant ∆POX01907 relative to that observed in the parental strain ∆PoxKu70 (Fig. 2b), which was consistent with our previous analyses.

To confirm the reduction in RSDE production in ∆POX01907 as being a result of POX01907 deletion, a complementary strain (CPOX01907) was constructed and confirmed by PCR (Additional file 5: Figure S2c) using specific primer pairs (Additional file 6: Table S3). Enzyme assays indicated no significant difference in RSDE production between CPOX01907 and ∆PoxKu70 (Fig. 2c).

Deletion of POX01907 promotes P. oxalicum mycelial growth during the late stage of SCS induction

Equal amounts of fresh spores collected from P. oxalicum ∆PoxKu70, the deletion mutant ∆POX01907, and the complementary strain CPOX01907 were inoculated on solid-medium plates in the presence of glucose, soluble corn starch as the sole carbon source, and potato dextrose agar (PDA), respectively, and cultured at 28 °C for 5 days. The results indicated that colony diameter of ∆POX01907 on all tested plates was significantly larger than that of ∆PoxKu70 and CPOX01907, which showed similar colony diameter (Fig. 3a). Moreover, the colony color of ∆POX01907 was lighter than that of ∆PoxKu70 and CPOX01907, which shared similar colony color (Fig. 3a).

Fig. 3
figure 3

Phenotype and growth comparison of ∆POX01907, the complementary strain CPOX01907, and the parental strain ∆PoxKu70 on different media. a Colonies on solid plates containing PDA, starch, and glucose, respectively, after incubation at 28 °C for 5 days. b, c Growth profiles of ∆POX01907 and ∆PoxKu70 in the presence of glucose and soluble corn starch, respectively. Data represent the means of three biological replicates

Additionally, we measured mycelial biomass in the three P. oxalicum strains grown in liquid media containing glucose or SCS. The mycelial weight of ∆POX01907 grown in glucose medium was similar to that of ∆PoxKu70 at 60-h post-inoculation and increased slightly after 60 h (Fig. 3b). The mycelial biomass of ∆POX01907 in SCS medium was slightly lower relative to that of ∆PoxKu70 at 36-h post-inoculation but increased significantly after 36 h (Fig. 3c).

POX01907 dynamically regulates the expression of major amylase genes in P. oxalicum

To investigate the regulatory roles of POX01907 in the expression of amylase genes in P. oxalicum, real-time quantitative reverse transcription PCR (RT-qPCR) was performed using total RNA from ∆PoxKu70 and ∆POX01907 grown on SCS medium for 4 h–48 h after a shift from glucose. Three major amylase genes were selected for evaluation, including an α-amylase gene POX09352/Amy13A, the RSDG gene POX01356/PoxGA15A, and a glucoamylase gene POX02412. The results revealed that transcript levels of POX01356/PoxGA15A and POX02412 in ∆POX01907 increased by 2752.1% and 506.0%, respectively, relative to those in the parental strain ∆PoxKu70 at 4 h post-SCS induction, whereas we did not observe changes in POX09352/Amy13A transcript levels. At 12 h, only POX01356/PoxGA15A continued to show elevation in transcript level in ∆POX01907 (by 212.3%), whereas POX09352/Amy13A transcript levels began to decrease (by 77.3%). The expression of all three genes was reduced by 71.3%–98.8% in ∆POX01907 after 12 h of SCS induction, although POX01356/PoxGA15A transcript level showed no significant difference from that in the ∆PoxKu70 at 24 h (Fig. 4).

Fig. 4
figure 4

Regulated expression of the major amylase genes by POX01907 in Penicillium oxalicum under the induction of soluble corn starch according to real-time quantitative reverse transcription polymerase chain reaction analysis. Transcript levels of amylase genes in the ∆POX01907 and ∆PoxKu70 investigated at 4, 12, 24 and 48 h after a shift from glucose to soluble corn starch. Expression levels were normalized against that in the parental strain ∆PoxKu70. Asterisks indicate significant differences (**P < 0.01) between ∆POX01907 and ∆PoxKu70, as assessed by Student’s t test

RNA-seq analyses reveal broad regulation of POX01907 in P. oxalicum

To analyze genome-wide regulation of POX01907 under SCS induction in P. oxalicum, RNA-seq was employed using total RNA collected upon a transfer of the deletion mutants ∆POX01907 and ∆PoxKu70 from glucose to SCS medium, followed by a 4-h incubation. In total, approximately 22 million of clean reads (length: 100 bp) for each sample were generated (Accession Number SRP116594), with > 90% of the clean reads mapped onto the genome of P. oxalicum HP7-1 wild-type strain (Additional file 1: Table S1). To assess the reliability of the generated transcriptome data, Pearson’s correlation coefficients (r) were calculated among three biological replicates for each sample, with results indicating high correlation (r > 0.95) (Additional file 7: Figure S3) and confirming their accuracy.

Gene expression was evaluated according to FPKM values calculated with the software package RSEM [22], and DEGs were screened with DESeq2 [23]. Comparative analysis of transcriptomes between ∆PoxKu70 and ∆POX01907 revealed 1003 DEGs, including 459 downregulated (–7.3 < log2 fold change < − 1.0) and 544 upregulated (1.0 < log2 fold change < 7.2) genes within the POX01907 regulon (Additional file 8: Table S5). KEGG annotation indicated that these DEGs were primarily involved in metabolism (72.9%), specifically carbohydrate metabolism (17.5%), amino acid metabolism (14.0%) and xenobiotic biodegradation and metabolism (10.0%) (Fig. 5a).

Fig. 5
figure 5

Transcriptome profiling of ∆POX01907 and the parental strain ∆PoxKu70 during growth in the presence of soluble corn starch as the carbon source. a Kyoto encyclopedia of genes and genomes annotation of proteins encoded by differentially expressed genes (DEGs) in ∆POX01907 relative to ∆PoxKu70. DEGs were identified according to a |log2 fold change| ≥ 1 and a P ≤ 0.05 as thresholds. b DEGs involved in glycolysis/gluconeogenesis and the TCA cycle. Upregulated DEGs are labeled in red, and downregulated DEGs are labeled in blue. c POX01907-specific regulation of DEGs encoding carbohydrate-active enzymes (CAZymes). Upregulated DEGs are labeled in pink, and downregulated DEGs are labeled in blue. Solid triangles represent DEGs encoding plant-cell-wall-degrading enzymes, and solid pentagrams represent DEGs encoding amylases. d Heatmap showing transcription levels of DEGs encoding putative transcription factors

Nutrients and energy required by P. oxalicum are derived from the substrate SCS. Of the 10 DEGs involved in starch and sucrose metabolism, five were involved in starch degradation. Among these, two glucoamylase genes (POX01356/PoxGA15A and POX02412) and two α-glucosidase genes (POX03889 and POX06242) showed increased expression (by 136.0–936.0%) in ∆POX01907 relative to the parental strain ∆PoxKu70, whereas expression of a gene encoding a 1,4-α-glucan-branching enzyme (POX04938) decreased by 50.7%. Surprisingly, seven genes were involved in cellulose degradation, including a cellobiohydrolase gene (POX05587/Cel7A-2), four endo-β-1,4-glucanase genes (POX05571/Cel7B, POX01206, POX07535/Cel12A, and POX06983), a β-glucosidase gene (POX07963), and a lytic polysaccharide monooxygenase gene (POX08897), with all displaying log2 fold changes ranging from −2.1 to 2.4. Additionally, seven genes were involved in glycolysis/gluconeogenesis and included a phosphoenolpyruvate carboxykinase gene (POX00835), a fructose-1,6-bisphosphatase gene (POX00441), two alcohol dehydrogenase genes (POX01238 and POX05829), two aldehyde dehydrogenase genes (POX02392 and POX03958), and an l-lactate dehydrogenase gene (POX09674), as well as two genes [pyruvate carboxylase (POX02290) and isocitrate dehydrogenase (POX08289)] involved in the citric acid cycle (TCA cycle). The transcript levels of these DEGs, except for POX02392 and POX03958, were upregulated in ∆POX01907 by 105.6%–255.3% (Fig. 5b).

In the regulon of POX01907, 96 DEGs were identified as encoding CAZymes, including 35 from the glycoside hydrolase family, 11 from the glycosyl transferase family, seven from the carbohydrate esterase family, eight from families exhibiting auxiliary activity, one polysaccharide lyase, and 11 from carbohydrate-binding-module families. Among these, 35 were downregulated (− 4.2 < log2 fold change < − 1.0) and 61 were upregulated (1.0 < log2 fold change < 4.5) (Fig. 5c). The regulon mainly contained five genes encoding starch-degrading enzymes previously described in this study, 20 genes encoding CWDEs, and seven genes predicted to encode enzymes degrading chitin. Notably, ∆POX01907 showed significant upregulation of genes encoding most of the CWDEs and amylase, and downregulation of chitin-degrading genes under SCS induction, suggesting multiple-regulation of genes involved in degrading different carbohydrates, including starch, cellulose, hemicellulose, and chitin.

Additionally, in the POX01907 regulon, 39 genes predicted as encoding TFs were detected, including 22 upregulated with a log2 fold change from 1.0 to 3.0 and 17 downregulated (− 3.2 < log2 fold change < − 1.0) (Fig. 5d). Functional annotation indicated that > 50% these contained zinc-related structures (Zn2Cys6 and C2H2), with two regulatory genes (POX01184 and POX04860/PDE_07199) reported to regulate cellulase gene expression in P. oxalicum [21]. However, no report was published about the regulation of these predicted TFs towards amylase genes expression.

POX01907 contains a pair of SANT/Myb domains

The POX01907 protein contains 1794 amino acids and two SANT [switching-defective protein 3 (Swi3), adenosine deaminase 2 (Ada2), nuclear receptor co-repressor (N-CoR), and TFIIIB)]/Myb domains [SANT/Myb833–881 (IPR001005; E-value: 1.38e−7) and SANT/Myb1086–1134 (IPR001005; E-value 3.03e−4)] (Fig. 6a). Additionally, BlastP analysis revealed that POX01907 shares 99% and 43% of identities with PDE_09981 in P. oxalicum 114-2 and AN8076.2 in A. nidulans FGSC A4 (XP_681345.1), respectively. The functions of PDE_09981 in P. oxalicum 114-2 and AN8076.2 in A. nidulans FGSC A4 (XP_681345.1) are unknown.

Fig. 6
figure 6

Characterization of POX01907 from Penicillium oxalicum HP7-1. a Modular architecture of POX01907. b Unrooted phylogenetic tree for POX01907 and its putative homologs. The phylogenetic tree was constructed using MEGA 7.0 software with the neighbor-joining method and a Poisson model. Bootstrap values shown were calculated with 1000 replicates

A phylogenetic tree for POX01907 and its homologs was constructed, revealing specificity to filamentous fungi, especially Penicillium spp. and Aspergillus spp. (Fig. 6b).

Discussion

In this study, we identified five novel regulatory genes (POX01907, POX03446, POX06509, POX07078, and POX09752) involved in mediating RSDE production in P. oxalicum through transcriptome profiling and genetic analysis. Further analyses confirmed that POX01907 regulated the expression of major amylase genes, including the RSDG gene PoxGA15A, as well as P. oxalicum mycelium growth in the presence of SCS. This represents the first report of POX01907 involvement in regulation of RSDG gene expression.

The SANT domain comprises an approximately 50-amino acid motif located in the subunits of many members of chromatin-remodeling complexes, such as Swi3, N-CoR, Ada2, and chromodomain-helicase-DNA binding protein 1 (Chd1), and consists of three α-helices arranged in a helix–turn–elix motif, with each α-helix containing a bulky aromatic residue and is similar to the Myb DNA-binding domain (DBD) [24]. However, the functions of SANT domains might be divergent from those of canonical Myb DBDs.

The SANT domain functions as a unique histone-interaction module that couples histone binding to enzyme catalysis and plays a central role in chromatin remodeling by regulating the activities of histone acetyltransferases and deacetylases to synergistically promote and maintain histone deacetylation [24]. SANT domains are capable of interacting with DNA (i.e., Saccharomyces cerevisiae Chd1 and Arabidopsis sp. Swc4 bind specific AT-rich DNA sequences in a non-canonical manner) [25, 26], histones [24], and other proteins (i.e., Chd1 interacts with the transcription-elongation factors Rtf1, Spt4–Spt5, and Spt6–Pob3) [27]. Additionally, the SANT protein Ada2 from Trichoderma reesei is required for mycelial growth, sporulation, and the expression of cellulase genes [28]. In filamentous fungi, including P. oxalicum, the transcription of cellulase and amylase genes is often co-regulated by several TFs, such as PoxNsdD [29], PoxAmyR [30], and PoxHmbB [15]. In the present study, comparative analysis of transcriptomes indicated that POX01907 also regulated the transcription of several cellulase genes in P. oxalicum under the induction of SCS. Therefore, we speculated that POX01907 might play an essential role in the expression of amylase genes by interacting with chromatin-remodeling complexes, although this requires further confirmation.

POX01907 dynamically regulated the expression of major RSDG and α-amylase genes similar to other known TFs identified previously in P. oxalicum [15, 21, 29, 30] and was dependent upon the nutrient and energy needs of fungal cells. Transcriptome profiling indicated that POX01907 had minimal influence on the expression of genes involved in the glycolysis pathway. During the early period of P. oxalicum cultivation, fungal cells require trace amounts of glucose for development and growth, and POX01907 inhibited the expression of glucoamylase genes POX01356/PoxGA15A and POX02402, thereby avoiding carbon catabolite repression. Along with glucose consumption and cell proliferation, POX01907 initiated the transcription of genes encoding glucoamylases and α-amylases, resulting in sufficient enzyme secretion to promote the degradation of starch into glucose.

Conclusions

Collectively, our results identified POX01907, a novel transcription factor gene responsible for regulating the production of RSDE through controlling the expression of the major RSDG gene POX01356/PoxGA15A, a glucoamylase gene (POX02412), and the α-amylase gene POX09352/Amy13A. These findings provide novel insights into the regulatory mechanism associated with fungal amylolytic enzymes production and their genes expression.

Methods

P. oxalicum strains and cultivation conditions

All P. oxalicum strains (Table 2) were cultured on PDA plates at 28 °C for 6 days to obtain fungal spores, resuspended using 0.1% Tween 80, and the concentration of fungal spores adjusted to 1.0 × 108/mL. P. oxalicum strains HP7-1 and ΔPoxKu70 were deposited in the China General Microbiological Culture Collection (Beijing, China) with Accession Numbers 10781 and 3.15650, respectively.

Table 2 Penicillium oxalicum strains used in this work

For measurement of RSDE production, P. oxalicum strains were cultivated according to previously described methods [5], with some modifications. Under non-transferring conditions, fresh spores (1.0 × 108/mL) of P. oxalicum strains were directly inoculated into minimal medium containing SCS as the sole carbon source at 28 °C for 4–6 days. Under transferring conditions, P. oxalicum spores (1.0 × 108/mL) were first inoculated into minimal medium containing glucose as the carbon source for 24 h, followed by transfer of an equal amount of mycelia from each P. oxalicum strain into minimal medium containing SCS as the carbon source for incubation at 28 °C for 2–4 days. For RNA-seq and RT-qPCR analyses, P. oxalicum strains were cultured for 4 h–48 h according to the methods described for transferring conditions.

Extraction of total DNA and RNA from P. oxalicum

Extraction of total DNA and RNA from P. oxalicum was performed as described previously [16]. Briefly, collected P. oxalicum mycelia were ground with liquid nitrogen, and lysate reagent [20 mM sodium acetate trihydrate, 10 mM ethylenediaminetetraacetic acid, 40 mM Tris–HCl, and 1% sodium dodecyl sulfate (pH 8.0)] was added and mixed. Phenol–chloroform was used to remove proteins. Total DNA was separated and collected by centrifugation at 11,300g for 10 min a Trizol RNA kit (Life Technologies, Carlsbad, CA, USA) was used to extract total RNA according to manufacturer instructions.

Construction of P. oxalicum gene-deletion mutants

Deletion of candidate genes from P. oxalicum was performed according to methods reported by Zhao et al. [16]. The knockout cassette for each candidate gene was constructed by fusion PCR and comprised a 1.9-kb G418-resistance gene and approximately 2 kb of the upstream and downstream DNA fragments of the target gene, which were amplified by PCR using the corresponding primer pairs (Additional file 6: Table S4). The generated knockout cassette was introduced into parental strain ∆PoxKu70 protoplasts, and selected transformants were further confirmed by PCR and/or Southern blot using specific primer pairs and/or probes (Additional file 6: Table S4).

Mutant complementation

A complementary strain of the deletion mutant ∆POX01907 was generated as described previously [21]. The complementary cassette comprised approximately 2 kb of the upstream- and downstream-flanking sequences of an aspartic protease gene (POX05007) used as the integrative locus in the genome, 1.2 kb of a DNA fragment encoding the bleomycin-resistance gene, and 7.8 kb of the complementary gene containing the promoter, coding region, and terminator. These four DNA fragments were amplified by PCR using specific primer pairs (Additional file 6: Table S4) and subsequently ligated together using a pEASY-ui seamless cloning and assembly kit (TransGen Biotech, Beijing, China). The generated complementary cassette was introduced into fresh ∆POX01907 protoplasts, and the resulting complementary strains were further confirmed by PCR.

Phenotypic investigation of P. oxalicum strains

Equal amounts of fresh spores from P. oxalicum strains, including the deletion mutant ∆POX01907, the complementary strain CPOX01907, and the parental strain ∆PoxKu70, were inoculated on solid plates containing glucose or SCS as the sole carbon source or PDA and incubated at 28 °C for 5 days. Colonies were photographed using a Canon EOS 6D digital camera (Canon, Beijing, China).

Biomass determination for P. oxalicum strains

Fresh 1.0 × 108 spores from P. oxalicum strains, including the parental strain ∆PoxKu70 and the deletion mutant ∆POX01907, were inoculated into 100 mL of glucose or starch liquid medium, respectively, and cultured at 28 °C for 12 h–72 h. The hypha was harvested using a vacuum filter every 12 h and dried to a constant weight at 50 °C.

RNA-seq analysis

RNA-seq analysis was performed according to methods described by Zhao et al. [16]. Total RNA extracted from P. oxalicum strains was used to construct cDNA libraries, with each cDNA having an average length of 100 bp. The constructed cDNA libraries were subjected to evaluation using an Agilent 2100 Bioanalyzer (Agilent Technologies, Santa Clara, CA, USA) and an ABI StepOnePlus real-time PCR system (Applied Biosystems, Forster City, CA, USA), and subsequently sequenced using an IIIumina HiSeq 4000 system (Illumina, San Diego, CA, USA). After quality control, the generated clean reads were mapped onto the P. oxalicum HP7-1 genome and functionally annotated using BWA v0.7.10-r789 (http://sourceforge.net/projects/bio-bwa/files/) and Bowtie2 v2.1.0 [31]. Gene-expression levels (FPKM) were calculated using RSEM v1.2.12 [22], and DEGs were screened and identified using NOISeq or DESeq2 [23]. Pearson’s correlation coefficient was used to evaluate transcriptome reliability among three biological replicates of each sample.

Southern hybridization analysis

Southern hybridization analysis of the deletion mutant ∆POX01907 and the parental strain ∆PoxKu70 was performed as previously described [16]. Briefly, total DNA of each strain was extracted and digested with PstI (TaKaRa, Dalian, China). After separation on an 0.8% agarose gel, the generated DNA fragments were transferred to a Hybond-N+ nylon membrane (GE Healthcare, Little Chalfont, UK). The probe used for Southern hybridization was amplified with the primers POX01907-probe-F and POX01907-probe-R (Additional file 4: Table S3). A DIG-High prime DNA labeling and detection starter kit (Life Technologies, Carlsbad, CA, USA) was used to investigate the hybridized bands.

RT-qPCR analysis

RT-qPCR was used to analyze differences in the expression levels of amylase genes between the deletion mutant ∆POX01907 and the parental strain ∆PoxKu70 according to a previously described method [16]. Total RNA from both ∆POX01907 and ∆PoxKu70 was extracted and used as a template to generate first-strand cDNA for RT-PCR using the PrimeScript RT regent kit with gDNA Eraser (TaKaRa). Each qPCR comprised a 20-μL volume, including 2.0 μL of the template for first-stand cDNA, 10 μL of SYBR Premix ExTaq II, 0.8 μL of 10 μM primer (either forward or reverse), and 6.4 μL of sterile water, subjected to initial denaturation for 3 min at 98 °C, followed by 40 cycles of 10 s at 98 °C and 30 s at 58 °C. Fluorescent signals were investigated at the end of each extension step at 80 °C according to the method described by Zhang et al. [10].

Enzyme activity and concentration

RSDE activity and concentration were measured as described previously [5]. Briefly, crude extract from P. oxalicum strains, including the deletion mutant ∆POX01907, the complementary strain CPOX01907, and the parental strain ∆PoxKu70, was added to 0.1 M citric acid/disodium hydrogen phosphate buffer (pH 4.5) containing 1.0% (w/v) raw cassava flour as the substrate, and the mixture was incubated at 65 °C for 30 min. Inactivated crude enzyme extract was used as the blank control. The generated reducing sugars were measured using the 3,5-dinitrosalicylic acid method [32] at 540 nm. One unit of enzymatic activity (U) was defined as the amount of enzyme required to produce 1 μmol of reducing sugars per min from the reaction substrate.

Intracellular enzyme concentration in P. oxalicum strains was measured using a Bradford assay kit (Pierce Biotechnology, Rockford, IL, USA) according to manufacturer instructions.

Phylogenetic analysis

POXO1907 homologs were downloaded from the NCBI (https://www.ncbi.nlm.nih.gov/) website following BlastP analyses (https://blast.ncbi.nlm.nih.gov/Blast.cgi?PAGE=Proteins). The phylogenetic tree was constructed using MEGA 7.0 software [33] with a neighbor-joining method and a Poisson correction model.

Statistical analysis

Microsoft Excel (Office 2016; Microsoft, Redmond, WA, USA) was used for the statistical analysis of all experimental data associated with enzyme production and gene transcription. Significance (P < 0.05 or P < 0.01) among samples was calculated using Student’s t test.

Accession numbers

All transcriptomic data are available from the SRA database (Accession Number SRP116594). DNA sequences of POX01907, POX03446, POX06509, POX07078 and POX09752 are available from the GenBank database (Accession Numbers MH742968-MH742972).

Abbreviations

CAZymes:

carbohydrate-active enzymes

CWDEs:

plant cell wall-degrading enzymes

DBD:

DNA-binding domain

DEGs:

differentially expressed genes

FPKM:

fragments per kilobase of exon per million mapped fragments

PDA:

potato-dextrose agar

RSDE:

raw-starch-digesting enzyme

RSDG:

raw-starch-digesting glucoamylase

SCS:

soluble corn starch

RT-qPCR:

real-time quantitative reverse transcription-PCR

TFs:

transcription factors

References

  1. Sun X, Liu Z, Qu Y, Li X. The effects of wheat bran composition on the production of biomass-hydrolyzing enzymes by Penicillium decumbens. Appl Biochem Biotechnol. 2008;146(1–3):119–28.

    Article  CAS  Google Scholar 

  2. Robertson GH, Wong DW, Lee CC, Wagschal K, Smith MR, Orts WJ. Native or raw starch digestion: a key step in energy efficient biorefining of grain. J Agric Food Chem. 2006;54:353–65.

    Article  CAS  Google Scholar 

  3. Rajoka MI, Yasmeen A. Induction, and production studies of a novel glucoamylase of Aspergillus niger. World J Microbiol Biotechnol. 2005;21:179–87.

    Article  CAS  Google Scholar 

  4. Hata Y, Ishida H, Kojima Y, Ichikawa E, Kawato A, Suginami K, Imayasu S. Comparison of two glucoamylases produced by Aspergillus oryzae in solid-state culture (koji) and in submerged culture. J Ferment Bioeng. 1997;84:532–7.

    Article  CAS  Google Scholar 

  5. Xu QS, Yan YS, Feng JX. Efficient hydrolysis of raw starch and ethanol fermentation: a novel raw starch-digesting glucoamylase from Penicillium oxalicum. Biotechnol Biofuels. 2016;9:216.

    Article  Google Scholar 

  6. Li HF, Chi ZM, Duan XH, Wang L, Sheng J, Wu LF. Glucoamylase production by the marine yeast Aureobasidium pullulans N13d and hydrolysis of potato starch granules by the enzyme. Process Biochem. 2007;42:462–5.

    Article  Google Scholar 

  7. Fujio Y, Morita H. Improved glucoamylase production by Rhizopus sp. A-11 using metal-ion supplemented liquid medium. J Ferment Bioeng. 1996;82:554–7.

    Article  CAS  Google Scholar 

  8. Ayodeji AO, Ogundolie FA, Bamidele OS, Kolawole AO, Ajele JO. Raw starch degrading, acidic-thermostable glucoamylase from Aspergillus fumigatus CFU-01: purification and characterization for biotechnological application. J Microbiol Biotechnol. 2017;6:90–100.

    Google Scholar 

  9. Lomthong T, Chotineeranat S, Kitpreechavanich V. Production and characterization of raw starch degrading enzyme from a newly isolated thermophilic filamentous bacterium, Laceyella sacchari LP175. Starch-Stärke. 2015;67:255–66.

    Article  CAS  Google Scholar 

  10. Zhang T, Zhao S, Liao LS, Li CX, Liao GY, Feng JX. Deletion of TpKu70 facilitates gene targeting in Talaromyces pinophilus and identification of TpAmyR involvement in amylase production. World J Microbiol Biotechnol. 2017;33:171.

    Article  Google Scholar 

  11. Xiong Y, Wu VW, Lubbe A, Qin L, Deng SW, Kennedy M, Bauer D, Singan VR, Barry K, Northen TR, Grigoriev IV, Glass NL. A fungal transcription factor essential for starch degradation affects integration of carbon and nitrogen metabolism. PLoS Genet. 2017;13(5):e1006737.

    Article  Google Scholar 

  12. Hu YB, Liu GD, Li ZH, Qin YQ, Qu YB, Song X. G protein-cAMP signaling pathway mediated by PGA3 plays different roles in regulating the expressions of amylases and cellulases in Penicillium decumbens. Fungal Genet Biol. 2013;58–59(5):62–70.

    Article  Google Scholar 

  13. Lei YF, Liu GD, Li ZH, Gao LW, Qin YQ, Qu YB. Functional characterization of protein kinase CK2 regulatory subunits regulating Penicillium oxalicum asexual development and hydrolytic enzyme production. Fungal Genet Biol. 2014;66(3):44–53.

    Article  CAS  Google Scholar 

  14. Chen L, Zou G, Zhang L, de Vries RP, Yan X, Zhang J, Liu R, Wang CS, Qu YB, Zhou ZH. The distinctive regulatory roles of PrtT in the cell metabolism of Penicillium oxalicum. Fungal Genet Biol. 2014;63(1):42–54.

    Article  CAS  Google Scholar 

  15. Xiong YR, Zhao S, Fu LH, Liao XZ, Li CX, Yan YS, Liao LS, Feng JX. Characterization of novel roles of a HMG-box protein PoxHmbB in biomass-degrading enzyme production by Penicillium oxalicum. Appl Microbiol Biotechnol. 2018;102:3739–53.

    Article  CAS  Google Scholar 

  16. Zhao S, Yan YS, He QP, Yang L, Yin X, Li CX, Mao LC, Liao LS, Huang JQ, Xie SB, Nong QD, Zhang Z, Jing L, Xiong YR, Duan CJ, Liu JL, Feng JX. Comparative genomic, transcriptomic and secretomic profiling of Penicillium oxalicum HP7-1 and its cellulase and xylanase hyper-producing mutant EU2106, and identification of two novel regulatory genes of cellulase and xylanase gene expression. Biotechnol Biofuels. 2016;9:203.

    Article  Google Scholar 

  17. Huber W, Carey VJ, Gentleman R, Anders S, Carlson M, Carvalho BS, Bravo HC, Davis S, Gatto L, Girke T, Gottardo R, Hahne F, Hansen KD, Irizarry RA, Lawrence M, Love MI, MacDonald J, Obenchain V, Oleś AK, Pagès H, Reyes A, Shannon P, Smyth GK, Tenenbaum D, Waldron L, Morgan M. Orchestrating high-throughput genomic analysis with bioconductor. Nat Methods. 2015;12:115–21.

    Article  CAS  Google Scholar 

  18. Lockington RA, Rodbourn L, Barnett S, Carter CJ, Kelly JM. Regulation by carbon and nitrogen sources of a family of cellulases in Aspergillus nidulans. Fungal Genet Biol. 2002;37:190–6.

    Article  CAS  Google Scholar 

  19. Breakspear A, Momany M. Aspergillus nidulans conidiation genes dewA, fluG, and stuA are differentially regulated in early vegetative growth. Eukaryot Cell. 2007;6(9):1697–700.

    Article  CAS  Google Scholar 

  20. Lee BY, Han SY, Choi HG, Kim JH, Han KH, Han DM. Screening of growth- or development-related genes by using genomic library with inducible promoter in Aspergillus nidulans. J Microbiol. 2005;43:523–8.

    CAS  PubMed  Google Scholar 

  21. Yan YS, Zhao S, Liao LS, He QP, Xiong YR, Wang L, Li CX, Feng JX. Transcriptomic profiling and genetic analyses reveal novel key regulators of cellulase and xylanase gene expression in Penicillium oxalicum. Biotechnol Biofuels. 2017;10:279.

    Article  Google Scholar 

  22. Li B, Dewey CN. RSEM: accurate transcript quantification from RNA-Seq data with or without a reference genome. BMC Bioinform. 2011;12:323.

    Article  CAS  Google Scholar 

  23. Love MI, Huber W, Anders S. Moderated estimation of fold change and dispersion for RNA seq data with DESeq2. Genome Biol. 2014;15:550.

    Article  Google Scholar 

  24. Boyer LA, Latek RR, Peterson CL. The SANT domain: a unique histone-tail-binding module? Nat Rev Mol Cell Biol. 2004;5:158–63.

    Article  CAS  Google Scholar 

  25. Gómez-Zambrano Á, Crevillén P, Franco-Zorrilla JM, López JA, Moreno-Romero J, Roszak P, Santos-González J, Jurado S, Vázquez J, Köhler C, Solano R, Piñeiro M, Jarillo JA. Arabidopsis SWC4 binds DNA and recruits the SWR1 complex to modulate histone H2A.Z. Mol Plant. 2018;11:815–32.

    Article  Google Scholar 

  26. Ryan DP, Sundaramoorthy R, Martin D, Singh V, Owenughes T. The DNA-binding domain of the Chd1 chromatin-remodelling enzyme contains SANT and SLIDE domains. EMBO J. 2011;30:2596–609.

    Article  CAS  Google Scholar 

  27. Simic R, Lindstrom D, Tran HG, Roinick KL, Costa PJ, Johnson AD, Hartzog GA, Arndt KM. Chromatin remodeling protein Chd1 interacts with transcription elongation factors and localizes to transcribed genes. EMBO J. 2003;22:1846–56.

    Article  CAS  Google Scholar 

  28. Zhang FL, Cao YL, Lv XX, Wang L, Li CY, Zhang WX, Chen GJ, Liu WF. A copper-responsive promoter replacement system to investigate gene functions in Trichoderma reesei: a case study in characterizing SAGA genes. Appl Microbiol Biotechnol. 2017;101:2067–78.

    Article  Google Scholar 

  29. He QP, Zhao S, Wang JX, Li CX, Yan YS, Wang L, Liao LS, Feng JX. Transcription factor PoxNsdD regulates the expression of genes involved in plant biomass-degrading enzymes, conidiation and pigment biosynthesis in Penicillium oxalicum. Appl Environ Microbiol. 2018;84:e01039-18.

    Article  Google Scholar 

  30. Li ZH, Yao G, Wu R, Gao L, Kan Q, Liu M, Yang P, Liu G, Qin Y, Song X, Zhong Y, Fang X, Qu Y. Synergistic and dose-controlled regulation of cellulase gene expression in Penicillium oxalicum. PLoS Genet. 2015;11:e1005509.

    Article  Google Scholar 

  31. Langmead B, Salzberg SL. Fast gapped-read alignment with Bowtie 2. Nat Methods. 2012;9:357–9.

    Article  CAS  Google Scholar 

  32. Miller GL. Use of dinitrosalicylic acid reagent for determination of reducing sugar. Anal Chem. 1959;31:426–8.

    Article  CAS  Google Scholar 

  33. Kumar S, Stecher G, Tamura K. MEGA7: molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol Biol Evol. 2016;33:1870–4.

    Article  CAS  Google Scholar 

Download references

Authors’ contributions

JXF designed, supervised this work, and was involved in the data analysis and manuscript preparation. SZ co-supervised all the experiments, wrote and revised the manuscript. MYZ carried out mutant construction, enzyme activity measurement, measurement of growth profiles, sample preparation for RNA-sequencing, and RT-qPCR. YNN performed RT-qPCR assay and was involved in the analysis of experimental data. CXL was involved in the bioinformatic analysis of the transcriptomes. LHF, RY, QW, CYW and HNX were involved in mutant construction and enzyme activity measurement. XML was involved in preparation of experimental materials and the analysis of experimental data. All authors read and approved the final manuscript.

Competing interests

The authors declare that they have no competing interests.

Availability of supporting data

All transcriptomic data are available from the SRA database (Accession Number SRP116594). DNA sequences of POX01907, POX03446, POX06509, POX07078 and POX09752 are available from the GenBank database (Accession Numbers MH742968-MH742972).

Consent for publication

All authors consent for publication.

Ethics approval and consent to participate

Not applicable.

Funding

This work was financially supported by grants from the Guangxi BaGui Scholars Program Foundation (Grant No. 2011A001), the Guangxi Natural Science Foundation (Grant No. 2012GXNSFGA060005), “One Hundred Person” Project of Guangxi and National Undergraduate Training Program for Innovation and Entrepreneurship.

Publisher’s Note

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

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Shuai Zhao or Jia-Xun Feng.

Additional files

Additional file 1: Table S1.

Summary of RNA-seq reads obtained for Penicillium oxalicum HP7-1.

Additional file 2: Figure S1.

Pearson’s correlation coefficient of the transcriptomes of Penicillium oxalicum HP7-1 among three biological replicates. P. oxalicum HP7-1 was cultivated in media containing glucose or soluble corn starch for 4 h after a shift from glucose.

Additional file 3: Table S2.

List of 916 differentially expressed genes in Penicillium oxalicum HP7-1 grown in the presence of starch as compared with that in the presence of glucose.

Additional file 4: Table S3.

List of 23 candidate regulatory genes determined in this study that regulate raw starch-degrading enzymes production in P. oxalicum HP7-1.

Additional file 5: Figure S2.

Confirmation of deletion of 11 candidate genes derived from ∆PoxKu70 and the complementary strain. a PCR analysis. M, 1-kb DNA marker; lanes 1–3, three transformants for each candidate gene; lane 4, ∆PoxKu70; and lane 5, ddH2O. The top panel shows amplification of the region to the left of the target gene, the middle panel shows amplification of the region to the right of the target gene, and the bottom panel shows amplification of the region of the target gene. b Southern hybridization analysis of the deletion mutant ∆POX01907. M, 1-kb DNA marker; lane 1, ∆PoxKu70; lane 2, ∆POX01907-7; lane 3, ∆POX01907-9; and lane 4, ∆POX01907-15. c PCR confirmation of the complementary strain CPOX01907. M, 1-kb DNA marker; lane 1, CPOX01907; lane 2, ∆PoxKu70; and lane 3, ddH2O. The top panel shows amplification of the bleomycin-resistance gene, and the bottom panel shows amplification of complementary cassette.

Additional file 6: Table S4.

Primers used in this study.

Additional file 7: Figure S3.

Pearson’s correlation analysis of the transcriptomes of Penicillium oxalicum strains ∆POX01907 and ∆PoxKu70 grown in medium containing soluble corn starch as the carbon source.

Additional file 8: Table S5.

List of 1003 genes differentially expressed in ∆POX01907 as compared with the parental strain ∆PoxKu70 grown on soluble corn starch as the sole carbon source.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhang, MY., Zhao, S., Ning, YN. et al. Identification of an essential regulator controlling the production of raw-starch-digesting glucoamylase in Penicillium oxalicum. Biotechnol Biofuels 12, 7 (2019). https://doi.org/10.1186/s13068-018-1345-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13068-018-1345-z

Keywords