The filamentous ascomycete fungus Aspergillus niger is a prolific secretor of organic acids, proteins, enzymes and secondary metabolites. Throughout the last century, biotechnologists have developed A. niger into a multipurpose cell factory with a product portfolio worth billions of dollars each year. Recent technological advances, from genome editing to other molecular and omics tools, promise to revolutionize our understanding of A. niger biology, ultimately to increase efficiency of existing industrial applications or even to make entirely new products. However, various challenges to this biotechnological vision, many several decades old, still limit applications of this fungus. These include an inability to tightly control A. niger growth for optimal productivity, and a lack of high-throughput cultivation conditions for mutant screening. In this mini-review, we summarize the current state-of-the-art for A. niger biotechnology with special focus on organic acids (citric acid, malic acid, gluconic acid and itaconic acid), secreted proteins and secondary metabolites, and discuss how new technological developments can be applied to comprehensively address a variety of old and persistent challenges.

For over a century, the mould Aspergillus niger has been gradually harnessed and modified into a multipurpose cell factory capable of converting a variety of cheap substrates into a range of valuable molecules (Figure 1 and Table 1). This global biotechnological revolution began in the 1920s, when Pfizer developed fermentation of citric acid from simple sugars using A. niger [1–3]. Around 20 years later, gluconic acid production using this fungus was also industrially successful [4,5]. Both these weak acids and their salts are used in a variety of applications as antioxidants, preservatives, acidulants, pH regulators or flavour enhancers in food, pharmaceutical and cosmetic industries. A. niger organic acid fermentation continues to grow annually, with citric and gluconic acid titres reported of 200 and 80 g/l, and markets predicted to reach 3.2 and 1 billion dollars by 2023 and 2027, respectively [6–9].

Technological and biological highlights from 100 years of A. niger biotechnology

Figure 1
Technological and biological highlights from 100 years of A. niger biotechnology
Figure 1
Technological and biological highlights from 100 years of A. niger biotechnology
Close modal
Table 1
Summary of key features for various A. niger isolates used in biotechnology
StrainIndustrial relevanceGenome accession numberGenome key findingsNHEJ mutant availableAssociated companiesReferences
CBS 513.88 Protein (glucoamylase) producer NCBI GCA_000002855 Seminal; over 14,000 genes predicted. Also identified putative secondary metabolite loci at genome level, RNA silencing pathways, and repeat induced point mutation MA70.15 MA169.4 DSM (Netherlands) [35,60,98
ATCC1015 Citric acid producer NCBI GCA_000230395 Gene duplication expanded genes necessary for the production of the citrate precursor oxaloacetate No JGI (USA) [99
SH2 Aconidial isolate and protein producer No genome sequence publicly available 11,517 predicted genes; strain is aconidial likely due to lack of prpA gene. Strain contains multiple SNPs in cell wall biosynthetic enzyme encoding genes No  [100
An76 Efficient lignocellulose degrader DNA Data Bank of Japan BCMY00000000 10,373 protein-coding genes, of which 79 are annotated to encode glycoside hydrolase No  [101,102
ATCC 10864 Biofilm forming isolate NCBI MCQH00000000 10,804 predicted genes, reason for biofilm formation phenotype unclear No  [103
WT-D Citric acid producer No genome sequence publicly available  D-10 Shanghai Industrial Microbiology Institute Tech. Co (China) [94
LDM3 Aconidial isolate and glucoamylase producer NCBI VTFN00000000 Non-synonymous mutations in 656 ORFs related to protein translation/ modification/secretion relative to CBS 513.88. SNP in tupA may cause aconidial phenotype No Longda Biotechnology (China) [104
JSC-093350089 International Space Station isolate NCBI MSJD00000000 Modified secondary metabolite profile relative to relative to ATCC1015 due to INDELs within promoter region of developmental regulator flbA CW12003 International Space Station/NASA (USA) [26
StrainIndustrial relevanceGenome accession numberGenome key findingsNHEJ mutant availableAssociated companiesReferences
CBS 513.88 Protein (glucoamylase) producer NCBI GCA_000002855 Seminal; over 14,000 genes predicted. Also identified putative secondary metabolite loci at genome level, RNA silencing pathways, and repeat induced point mutation MA70.15 MA169.4 DSM (Netherlands) [35,60,98
ATCC1015 Citric acid producer NCBI GCA_000230395 Gene duplication expanded genes necessary for the production of the citrate precursor oxaloacetate No JGI (USA) [99
SH2 Aconidial isolate and protein producer No genome sequence publicly available 11,517 predicted genes; strain is aconidial likely due to lack of prpA gene. Strain contains multiple SNPs in cell wall biosynthetic enzyme encoding genes No  [100
An76 Efficient lignocellulose degrader DNA Data Bank of Japan BCMY00000000 10,373 protein-coding genes, of which 79 are annotated to encode glycoside hydrolase No  [101,102
ATCC 10864 Biofilm forming isolate NCBI MCQH00000000 10,804 predicted genes, reason for biofilm formation phenotype unclear No  [103
WT-D Citric acid producer No genome sequence publicly available  D-10 Shanghai Industrial Microbiology Institute Tech. Co (China) [94
LDM3 Aconidial isolate and glucoamylase producer NCBI VTFN00000000 Non-synonymous mutations in 656 ORFs related to protein translation/ modification/secretion relative to CBS 513.88. SNP in tupA may cause aconidial phenotype No Longda Biotechnology (China) [104
JSC-093350089 International Space Station isolate NCBI MSJD00000000 Modified secondary metabolite profile relative to relative to ATCC1015 due to INDELs within promoter region of developmental regulator flbA CW12003 International Space Station/NASA (USA) [26

Note that a total of 17 genomes are available [10]. Abbreviations: NHEJ, non-homologous end-joining (used to increase targeting of exogenous DNA in recipient genome usually via kusA deletion or disruption, see main text); SNP: single nucleotide polymorphism; INDEL: insertion or deletion.

Other useful organic acids may soon be produced by A. niger at an industrial scale. Malic acid, for example, is currently used in food and pharmaceutical industries, and is a potentially sustainable replacement for the commodity chemical maleic anhydride [10]. Although malic acid is mainly produced chemically from fossil derived substrates, recent studies have engineered enhanced titres in submerged A. niger fermentation (e.g., 200 g/l), suggesting bioproduction of this product may soon be economically feasible [11–13]. Additionally, new acids have been added to the A. niger product portfolio, as evidenced by engineering itaconic acid producing isolates by heterologous expression of A. terreus biosynthetic and transport genes [14].

The other major application of A. niger is enzyme production, which broadly began in the late 1950s when advances in chromatography purification technology enabled biotechnologists to study a diverse repertoire of useful proteins. This included lactoferrin, lipase, arabinase, asparaginase, catalase, cellulase, β-galactosidase, glucoamylase, pectinase, phytase, proteases, xylanase, glucose oxidase and hemicellulase, amongst others [15]. The most successful application of these enzymes is probably glucoamylase saccharification of starch to glucose, a technology used by several multinational biotechnological companies worth over a billion dollars per year [16]. Fermentation of the glucoamylase enzyme GlaA in A. niger reaches titres of 30 g/l [17].

In the last decades, A. niger has also emerged as a producer of secondary metabolites (SMs) with titres of several g/l [17,18]. For example, heterologous expression of a Fusarium oxysporum non-ribosomal peptide synthetase (NRPS) encoding gene produced g/l titres of enniatin [19], which was previously developed into the antimicrobial drug fusafungine [20]. Excitingly, by modifying growth media amino acids, new-to-nature enniatin variants could be produced [19]. Additionally, by domain swap experiments in the NRPS encoding gene, production of hybrid non-ribosomal peptides which showed enhanced antimicrobial activity was possible [21].

Most recently, a new frontier in A. niger biotechnology has emerged, as applications of this fungus have looked to earth’s orbit and beyond. A. niger is one of the few fungi known to grow robustly inside the international space station [22–25], and metabolite profiling suggests astronautic isolates generate elevated levels of the antioxidant pyranonigrin A, which may enhance protection to radiation [26]. It is therefore hypothesized that A. niger is an outstanding candidate for use in a variety of space biotechnological applications, from generating useful products during extended space flight to terraforming efforts [25].

A huge investment of A. niger research has determined precise abiotic growth conditions necessary for optimal product titres, ranging from pH, carbon source, spore inoculum density, oxygenation, among many other parameters (for a recent review, see [27]). More recently, a combination of molecular, cellular and omics approaches has been used to comprehensively understand the molecular and cellular mechanisms of product biosynthesis and secretion/export.

Organic acid synthesis and export

Most industrially relevant organic acids produced by A. niger are metabolites of the tricarboxylic acid cycle (TCA), which are produced under specific growth stages and in response to various abiotic cues (e.g., phosphate limitation, low pH [28,29]). One notable acid that is broadly independent from the TCA cycle is gluconic acid, which is biosynthesized from a glucose substrate by the extracellular, cell wall-localized glucose oxidase GoxC [30].

Citric acid is formed from precursors oxaloacetate, acetyl-CoA and water at the final stage of the TCA by the mitochondrial citric acid synthase CitA [31]. Citric acid that is not recycled back into the TCA cycle can be transported to the cytosol by malate/citrate shuttles in the mitochondrial membrane including CtpA [32], which can then be exported from the cell by the recently discovered putative major facilitator superfamily transporter CexA [33]. Precisely why A. niger secretes citric acid into the external media is unknown, but this may occur after a diauxic shift to phosphate limited growth [28], and to increase iron bioavailability [29]. Genetic engineering of genes encoding TCA enzymes and substrate transporters have reported varied success in citric acid biotechnological applications. For example, over-expression of citA [31] and other TCA pathway enzymes [34] does not increase titres of citric acid, and disruption of ctpA does not impact productivity [32]. In contrast, over-expression of the CexA transporter encoding genes using either the inducible Tet-on system [35,36] or high constitutive expression increases citric acid titres by approximately 5 and 3 times, respectively [33].

Fermentation of the emerging value product malic acid is highly interconnected with that of citric acid due to the role of malic acid the TCA cycle. Additionally, as noted above, cytosolic malic acid is thought to be imported into the mitochondria by the malate/citrate shuttle CtpA. Increase of cytosolic malic acid has therefore been hypothesized to initiate citric acid fermentation in A. niger due to the exchange of malic acid and citric acid across the mitochondrial membrane [37]. However, it should be noted that ctpA mutants display comparable levels of both citric and malic acid in culture filtrates compared with progenitor controls – possibly due to functional redundancy of this transporter – thus limiting applications for genetic engineering of CtpA [32].

Despite these challenges, malic acid titres can be successfully increased by genetic engineering of A. niger membrane transporters, in a comparable approach to that of CexA and citric acid. The major A. niger malic acid exporter was identified by homology searches using the sequence from the fission yeast malate/succinate proton symporter Mae1 [12]. By overexpression of the A. niger orthologous gene, named dctA, malic acid titres could be increased by over 20% compared with controls [12]. Elsewhere, it was recently demonstrated that engineering cexA can be used to hyperproduce malic acid [11]. Disruption of cexA concomitantly reduces titres of citric and malic acid, with the mutant having reduced expression of genes for glucose transportation and the glycolytic pathway [11]. By overexpressing key genes in the cexA disruption mutant, including a low affinity glucose membrane transporter mstC, hexokinase hxkA and pyruvate kinase pkiA, among other genes, malic acid titres could be increased by ∼23% compared with the progenitor control [11]. Taken together, recent functional characterizations of CexA and other membrane transporters can be viewed as breakthroughs in A. niger biotechnology, which may soon be applied to generate organic acid hyper-producing isolates at an industrial level, especially when combined with other recent genetic strategies for maximizing productivity [13,38]. More generally, discovery and characterization of all A. niger organic acid organelle and plasma membrane transporters, and their detailed mechanistic functions, are crucial future objective for biotechnologists.

It is also notable that A. niger is a potential chassis host for production of itaconic acid, an emerging bulk chemical commercially produced by A. terreus [39]. In this species, cis-aconitate from the TCA cycle is transported out of mitochondria by the transporter MttA, after which it is converted to itaconate by the cis-aconitate decarboxylase CadA [14]. Ultimately, itaconate is exported from A. terreus by the membrane transporter mfsA [14]. A. niger lacks a known cadA orthologue, and expression of a codon optimized cadA and mttA was sufficient for proof of principle production of itaconic acid, with export from the cell presumably performed by a native membrane transporter [14]. By overexpressing the A. niger cytosolic citric acid synthase citB [40], and by identification and deletion of itaconic acid bioconversion pathways [41], titres above 30 g/l could be achieved in subsequent studies, which may enable the use of this molecule as a sustainably produced bulk chemical.

Protein secretion

The classical protein secretion route involves trafficking of cargo-loaded vesicles along microtubules and actin cables from the endoplasmic reticulum (ER)/Golgi to the hyphal tip [42–44]. Vesicles cluster at a sub-apical site called the Spitzenkörper, and become fused with the plasma membrane by a multiprotein complex termed the exocyst [45], after which cargo is released. In addition to extracellular enzymes for nutrient hydrolysis, cell wall synthesizing proteins are trafficked in vesicles, hence coupling growth and secretion at the hyphal tip. Exocytic addition of vesicle membrane to the plasma membrane is balanced by endocytosis, most likely at a sub-apical actin ring, which is important to maintain hyphal polarity [46]. The majority of industrially important enzymes probably follow this route, as evidenced by N-terminal secretion signals, and several studies that localize enzymes such as the glucoamylase GlaA to secretory vesicles (46–48). Modulation of classical secretion, for example by titratable expression of GTPase regulators of vesicle trafficking [46], or by constructing mutants with branching defects and elevated hyphal tip numbers [47,49] is an effective method for elevating secretion of both total protein and enzymes such as GlaA.

In addition to vesicle trafficking to the hyphal tip, other non-classical routes certainly play a role in A. niger protein secretion. There is growing evidence that proteins may exit hyphae at intercalary regions, most notably septal junctions [46,48,50–52], which could conceivably be an extracellular route for industrially relevant proteins, including GlaA [46]. Additionally, several proteins that lack any secretion signal for translocation to the ER are found in culture supernatants [53]. The A. niger aspartic protease PepN is secreted by non-classical routes, which are independent of autophagic processes necessary for unconventional secretion in Saccharomyces cerevisiae [54]. Excitingly, extracellular vesicles from a member of the fungal kingdom were first characterized in Cryptococcus neoformans [55] despite their probable observation in A. nidulans 40 years earlier [56]. Extracellular vesicles thus might exist in A. niger. In other fungi they can contain proteins, nucleic acids, carbohydrates, and lipids (reviewed in [57]), and could thus be relevant for industrial applications. Currently, mechanistic understanding of these unconventional secretion pathways is limited, and discoveries in this area may enable new avenues for elevating titres of proteins and other products during fermentation.

Secondary metabolites

Most A. niger SMs are produced during low or zero growth [58] by biosynthetic gene clusters (BGCs) consisting of one or more key enzyme encoding genes, such as a polyketide synthase (PKS) or NRPS [59]. These enzymes are modular, and function respectively by loading acetyl-CoA or an amino acid substrate onto a carrier protein. Next, the molecule is elongated and chemically modified by various modules, after which a nascent SM is released from the carrier protein through activity of a thioesterase domain [59]. Enzymes normally encoded by contiguous genes within the BGC then modify of the core moiety, for example by glycosylation, acylation or methylation. The BGC may also encode membrane transporters necessary for internal transport/export of the SM, and transcription factors which regulate BGC expression.

The first A. niger genome sequence was published in 2007 [60], with 78 predicted BGCs putatively involved in SM biosynthesis [61], the highest in any Aspergillus species. The majority of these BGCs produce an unknown product; recent data mining from 283 microarray experiments under a diverse range of experimental conditions suggests that most are either transcriptionally silent in the lab, or expressed in strictly limited circumstances [62]. Consequently, the potential for experimentally activating A. niger BGCs by specialized (co)cultivation conditions are promising for new bioactive molecule discovery [63]. Molecular approaches for SM activation have also been used in other filamentous fungi, including overexpression of the NRPS/PKS/BGC transcription factor, or through modification of epigenetic regulators which modulate BCG expression by chromatin remodelling [64].

Biosynthesis of most SMs is speculated to be highly compartmentalized and spatially organized in the cell. For example, penicillin synthesis in Penicillium chrysogenum (which can be considered one of the best characterized SM biosynthetic pathways) includes storage of amino acid precursors in vacuoles, biosynthesis both in the cytosol and peroxisomes, and secretion by either a hypothetical membrane transporter or in secretory vesicles (for a recent review see [65]). In A. nidulans, which also produces penicillin, targeting of the cytosolic NRPS AcvA to peroxisomes significantly increased titres [66], highlighting how understanding and modulating biosynthetic enzyme subcellular localization can be an effective future strategy for A. niger.

Connecting three product groups

Traditionally, A. niger biotechnological research has been separated into protein, organic acid and SM sub-disciplines. There is however a recent trend to holistically study these related fields and to understand primary metabolism including redox and cofactor metabolism as foundational basis for all products formed [1,3,67]. One recent discovery that has connected all three disciplines began by investigating a non-acidification phenotype in the UV mutagenized A. niger isolate D15 [68]. Non-acidification of growth media is of major interest for enzyme production isolates, as secretion of acid-induced proteases will be inhibited [69].

Accordingly, bulk segregant analysis was applied to mutant D15, a powerful technique which is broadly applicable to identify single-nucleotide polymorphisms (SNPs) in many mutagenized A. niger strains (Figure 1, [68,70]). In this approach, parasexual crosses are induced on solid medium between the strain of interest and a non-mutagenized control by selection and subsequent counter selection of heterokaryotic mycelia. Segregant progeny displaying the phenotype of interest – such as non-acidification of growth media – then have DNA pooled and bulk sequenced at sufficient depth so that multiple SNPs can be identified [70]. Ultimately, the SNP that causes the phenotype of interest is conserved in all the progeny, and absent in the control isolate.

Intriguingly, this approach demonstrated that non-acidification by A. niger mutant D15 was due to a SNP in the putative methyltransferase gene encoding a master regulator of secondary metabolism LaeA [68]. A follow up study demonstrated that LaeA regulates acidification by histone H3K4 and H3K9 methylation in the cexA promoter of the shochu producer A. luchuensis mut. kawachii [71], thus likely mechanistically explaining the non-acidification of A. niger D15 [68]. From SM perspectives, deletion and overexpression approaches have confirmed the role of A. niger LaeA in regulating hundreds of A. niger SM genes [72]. Finally, another most recent study linked for the first time the intracellular redox status of A. niger, which is dependent from the sugar and organic acid metabolism, with protein and glucoamylase production. It was demonstrated that metabolic engineering through NADPH cofactor engineering can indeed cause a substantial increase in GlaA titers [67].

Many more insights into the interconnections between organic acid, protein and secondary metabolite production are likely to be revealed in future studies, which may enable rational design of hyperproduction strains. Despite the above advances, several challenges exist for (re)engineering A. niger as a maximally efficient cell factory. Two key challenges are discussed below.

The control of filamentous morphology in an industrial A. niger application is crucial for any kind of submerged production process and can happen on two basic levels, micromorphology and macromorphology, describing the structure of individual hyphae and the characteristics of an entire aggregate of mycelium, respectively (reviewed in [27]). Whereas the beneficial impact of altering micromorphology of A. niger has been established in some instances (e.g., a hyperbranching phenotype on protein productivity [27]), a significant challenge is precise, user defined marcomorphological measurement and control in submerged culture. Macromorphology of A. niger in submerged culture can be divided into three classes: dispersed mycelium (free, individual hyphae), clumps (groups of hyphae loosely attached, forming undefined aggregates) and pellets (dense, spherically symmetric hyphal aggregates with a spore centre, originating from co-aggulative pellet forming mechanism) [73,74]. While dispersed cultures show a higher viscosity, thus lower mixability and oxygen transfer in the culture broth, the nutrient supply stays uniformly stable throughout the entire population. In a pellet forming process on the other hand, the mixability and oxygen transfer in the culture broth are higher, but the densely packed hyphal structures provoke oxygen and nutrient gradients within single pellets, thus causing limitations inside the pellet core [27]. Due to the often described, yet not fully understood correlation between growth/morphology and productivity in A. niger, both morphological forms can be favourable, depending on the desired product [75,76]. As classical, vesicle-based protein secretion and growth are strictly coupled processes, the well nutrient-supplied, thus fast-growing dispersed hyphae as well as the outer, non-nutrient limited regions of pellets are assumed to be ideal production hosts. For the production of secondary metabolites, which peaks at zero or very low growth, the inactive cores of pellets are probably well suited [77].

The predominant macromorphology present in a culture is classically controlled by process parameters like inoculum, medium composition, pH, temperature or mechanical forces [78]. In addition to this, novel approaches in the field of genetic engineering allow the titration of macromorphology without changing process conditions. It was for example shown recently, that the reduced expression of the putative gamma-adaptin encoding gene (aplD), results in a hyperbranching phenotype with defects in pellet formation, and could thus give a possible target for morphology engineering [75].

While dispersed mycelia can be analysed comparatively easily using light microscopy, this technique is not sufficient to image the complex, 3D structure of filamentous pellets, and especially their cores. To overcome this problem, pellet-slicing and subsequent light microscopic investigation has become a common tool [79]. The obvious disadvantage of this procedure is the destructive nature of slicing. Other approaches to examine pellets include confocal laser scanning microscopy of previously frozen and died slices, allowing to gain 3D insights into parts of the pellet's structure [80] as well as scanning electron microscopy of entire pellets, unveiling the difference between highly intertwined superficial hyphae and a densely packed mycelium [81]. In addition to these imaging-based methods, flow cytometry can be used to get information about the core compactness of fungal pellets [82]. However, none of these methods allow a non-destructive, detailed morphological analysis of an entire pellet. This problem was recently addressed with X-ray microtomography (µCT) [83]. Here, entire freeze-dried fungal pellets were investigated using X-ray and computer-based image analysis techniques, giving the exact distribution of hyphae and the number of hyphal tips and branches inside a pellet. The key advantage of this novel µCT technique is thus the possibility to determine the exact inner structure of intact fungal pellets and to calculate the effective diffusion factor inside an A. niger pellet in relation to the hyphal fraction [84]. Remarkably, it was shown that the correlation between effective diffusion factor and hyphal fraction in fungal pellets follows a universal law, which is independent of micromorphological features like hyphal diameter, branching angle or interval and growth angle [85]. With this new tool, it is conceivable to calculate the actual nutrient or oxygen gradient within a pellet in a foreseeable future. To do so, two major elements must be developed; a term describing the uptake or production rate inside the pellet and a second term for the diffusive mass transport through the pellet. As soon as both these elements come together, a detailed distribution of each individual substance inside fungal pellets can be predicted.

By combining the techniques described here, hopefully it will shortly be possible to design optimized macromorphology in each individual A. niger process as follows: (i) depending on the desired product, a suitable overall macromorphology is selected (dispersed/pellet). If the chosen form is pelleted, optimization is needed: (ii) nutrient gradient calculation to reverse engineer a theoretical pellet structure, optimized in nutrient and oxygen supply for the production of the specific substance; (iii) genetic and process engineering techniques to adjust the macromorphology to the calculated optimum; (iv) X-ray microtomography or other methods to examine the resulting macromorphological structures. Consequently, morphology engineering of A. niger chassis isolates with a pre-programmed macromorphology in submerged culture will drastically reduce screening time for putative hyperproduction mutants (Figure 2).

Two key technological developments have drastically increased the throughput of A. niger null, overexpression and conditional-expression mutant construction. Firstly, deletion of non-homologous end joining components significantly elevate the rates of homologous recombination between exogenous DNA cassettes and the A. niger genome, thus increasing targeting of desired sequences at user define loci [86]. In A. niger, this is conventionally achieved by deletion or disruption of the kusA gene, which encodes a subunit of the Ku heterodimer [87]. Such kusA mutants are available in organic acid and protein industrial strain lineages (Table 1). Secondly, genome editing using Cas9, Cas12a and other guided nucleases now enable rapid and precise deletion, disruption and insertion of desired sequences throughout the genome [88–90]. Consequently, it is now conceivable for single studies to generate libraries of several dozen or even hundreds of mutants with putatively elevated product titres.

Despite these advances, a significant bottleneck to screening such A. niger libraries for improved productivity is the limited throughput of shake flask and bioreactor based submerged culture, which are limited to several dozen technical replicates per user/week (Figure 2). Microtiter assays of liquid growth have been widely used in instances where macromorphological development is broadly irrelevant, for example high-throughput measurements of antifungal efficacy against A. niger [91]. However, it is currently unclear if it is possible to meaningfully replicate macromorphological changes and productivity observed in bioreactors (L) or shake flasks (ml) in smaller volumes (µl). Encouragingly, there has been success in culture miniaturization using other filamentous fungal cell factories. For example, a microtitre plate assay for screening dozens of A. vadensis transformants expressing an A. niger feruloyl esterase B, an enzyme required for dissociation of lignin from hemicellulose was developed [92]. High-throughput screens of submerged growth enabled clones with undetectable or minimal levels of overexpressed protein to be rapidly identified. Importantly, this study demonstrated that enzyme activity of clones was qualitatively comparable at either shake flask or microtiter scales. However, the authors observed an absolute decrease in enzyme activity around tenfold between shake flask and microtiter cultivations [92]. Therefore, despite the obvious advantages in throughput using this approach, microtiter cultivation may not be appropriate where detection of the desired product is challenging. Others were able to reproducibly ferment approximately 80g/l itaconic acid using A. terreus bioreactor (400 ml), shake flask (100 ml) and 96 well (100 µl) cultivations [93]. As the authors noted, the small size of A. terreus pellets (∼100 µm diameter) probably enabled comparable productivity across cultivation platforms. Are such scale down methods possible for A. niger, where pellets in shake flask/bioreactor models of either protein or citric acid fermentation reach over several mm in diameter [75,94]?

Current and future A. niger strain engineering cycles

Figure 2
Current and future A. niger strain engineering cycles

The cycle begins with a candidate gene list(s) for functional analysis derived from previous datasets/experiments. Numbers in parentheses denote either the total amount of time for each step (months per scientist, M) or an exemplar number of genes remaining in each step (e.g., n = 50 candidate genes). Refinement of functional predictions is now possible by publicly available coexpression resources for A. niger. Mutants can be generated in-house using kusA mutants/genome editing technology, or in future from acquiring isolates from a mutant library resource (dotted arrow). Next, preliminary fermentation can be conducted, currently in shake flask, but in future via MTP cultivation. Putative hyperproducers (which in this example has generated 1-5 isolates) can then be tested via extensive design of experiment (DoE) approaches or, in future, by functional analysis in pre-generated chassis strains with a defined macromorphology. Pilot production strains can then be further analysed for another iteration of strain engineering, or taken for scale up during trial industrial production. This engineering cycle is similar for other industrially used filamentous fungi, both with regards to the main steps, time-frames, and major bottle-necks.

Figure 2
Current and future A. niger strain engineering cycles

The cycle begins with a candidate gene list(s) for functional analysis derived from previous datasets/experiments. Numbers in parentheses denote either the total amount of time for each step (months per scientist, M) or an exemplar number of genes remaining in each step (e.g., n = 50 candidate genes). Refinement of functional predictions is now possible by publicly available coexpression resources for A. niger. Mutants can be generated in-house using kusA mutants/genome editing technology, or in future from acquiring isolates from a mutant library resource (dotted arrow). Next, preliminary fermentation can be conducted, currently in shake flask, but in future via MTP cultivation. Putative hyperproducers (which in this example has generated 1-5 isolates) can then be tested via extensive design of experiment (DoE) approaches or, in future, by functional analysis in pre-generated chassis strains with a defined macromorphology. Pilot production strains can then be further analysed for another iteration of strain engineering, or taken for scale up during trial industrial production. This engineering cycle is similar for other industrially used filamentous fungi, both with regards to the main steps, time-frames, and major bottle-necks.

Close modal

An innovative approach combines microparticle based reduction in A. niger pellet size with growth in microtiter wells [95]. By modifying inoculum and microparticle density, it is possible to both elevate production of a food flavouring compound in this system, while reproducibly generating pellets of uniform size [95]. In another study, microtiter cultivation of A. giganteus was combined with online measurements of fungal growth using scattered light in the BioLector system [96]. This approach also relied on reducing pellet formation, specifically by supplementing growth media with CaCl2 [96]. This approach was further developed into an automated workflow for BioLector-assisted microtiter plate cultivation of A. carbonarius that included an automated pipeline of sample processing, injection, image acquisition and morphology analysis [97]. These studies therefore demonstrate that reproducible growth in microtiter format of pellet forming fungi is indeed possible; however this currently requires modification of culture conditions to inhibit formation of large pellets. Therefore, while biotechnologists can currently apply high-throughput microtiter cultivations in A. niger, a limitation to this approach is that certain products which are favourably or exclusively secreted by large pellets will not be detectable [27]. Addressing this bottleneck will drastically increase rational strain engineering throughput in the future (Figure 2).

  • A. niger is a microbial cell factory used to produce proteins, organic acids, and secondary metabolites.

  • The molecular, cellular and metabolic basis of product biosynthesis and secretion are currently being revealed, suggesting they are highly interconnected.

  • The existing A. niger toolkit and available resources promise a rapid period of genetic, metabolic, and morphology engineering for efficient cell factories.

The authors declare that there are no competing interests associated with the manuscript.

The authors thank the Deutsche Forschungsgemeinschaft (DFG) for the financial support for this work [grant ME 2041/5‐1 within the SPP 1934 DiSPBiotech and grant ME 2041/13-1]. The authors thank the DFG for open access funding.

T.C., L.B. and V.M. co-wrote the text and generated the figures.

None

BGC

biosynthetic gene cluster

INDEL

insertion or deletion

NHEJ

non-homologous end-joining

NRPS

non-ribosomal peptide synthetase

PKS

polyketide synthase

SM

secondary metabolite

SNP

single nucleotide polymorphism

1.
Meyer
V.
,
Basenko
E.Y.
,
Benz
J.P.
,
Braus
G.H.
,
Caddick
M.X.
,
Csukai
M.
et al.
(
2020
)
Growing a circular economy with fungal biotechnology: a white paper
.
Fungal Biol. Biotechnol.
7
,
5
[PubMed]
2.
Currie
J.N.
(
1917
)
The citric acid fermentation of Aspergillus niger
.
J. Biol. Chem.
31
,
15
37
3.
Meyer
V.
,
Andersen
M.R.
,
Brakhage
A.A.
,
Braus
G.H.
,
Caddick
M.X.
,
Cairns
C.T.
et al.
(
2016
)
Current challenges of research on filamentous fungi in relation to human welfare and a sustainable bio-economy: a white paper
.
Fungal Biol. Biotechnol.
3
,
1
17
[PubMed]
4.
Porges
N.
,
Clark
T.F.
and
Gastrock
E.A.
(
1940
)
Gluconic acid production
.
Ind. Eng. Chem.
33
,
1065
1067
,
8
5.
Blom
R.H.
,
Pfeifer
V.F.
,
Moyer
A.J.
,
TraufRer
D.H.
,
Conway
H.F.
,
Crocker
C.K.
et al.
(
1952
)
Sodium gluconate production. Fermentation with Aspergillus niger
.
Ind. Eng. Chem.
44
,
435
440
,
2
6.
(
2020
)
Global Citric Acid Market Top Countries Data 2020: Industry Size, Share, Future Challenges, Revenue, Demand, Industry Growth and Top Players Analysis and forecast By 360 Market Updates
.
7.
Gluconic Acid Market Segmentation By Component (Glucono Delta-Lactone, Gluconic Acid, Calcium Salt Of Gluconic Acid, Sodium Salt Of Gluconic Acid and Iron Slat Of Gluconic Acid); By End-User Industry (Food And Beverage, Pharmaceutical, Chemicals, Agricult
.
[cited 2020 Oct 14]. Available from
8.
Tong
Z.
,
Zheng
X.
,
Tong
Y.
,
Shi
Y.-C.
and
Sun
J.
(
2019
)
Systems metabolic engineering for citric acid production by Aspergillus niger in the post-genomic era
.
Microb. Cell Fact
18
,
28
[PubMed]
9.
Zhang
H.
,
Zhang
J.
and
Bao
J.
(
2016
)
High titer gluconic acid fermentation by Aspergillus niger from dry dilute acid pretreated corn stover without detoxification
.
Bioresour. Technol.
203
,
211
219
[PubMed]
10.
Kövilein
A.
,
Kubisch
C.
,
Cai
L.
and
Ochsenreither
K.
(
2020
)
Malic acid production from renewables: a review
.
J. Chem. Technol. Biotechnol.
95
,
513
526
,
3
11.
Xu
Y.
,
Zhou
Y.
,
Cao
W.
and
Liu
H.
(
2020
)
Improved production of malic acid in Aspergillus niger by abolishing citric acid accumulation and enhancing glycolytic flux
.
ACS Synth. Biol.
9
,
1418
1425
,
6
[PubMed]
12.
Cao
W.
,
Yan
L.
,
Li
M.
,
Liu
X.
,
Xu
Y.
,
Xie
Z.
et al.
(
2020
)
Identification and engineering a C4-dicarboxylate transporter for improvement of malic acid production in Aspergillus niger
.
Appl. Microbiol. Biotechnol.
104
,
9773
9783
[PubMed]
13.
Xu
Y.
,
Shan
L.
,
Zhou
Y.
,
Xie
Z.
,
Ball
A.S.
,
Cao
W.
et al.
(
2019
)
Development of a Cre-loxP-based genetic system in Aspergillus niger ATCC1015 and its application to construction of efficient organic acid-producing cell factories
.
Appl. Microbiol. Biotechnol.
103
,
8105
8114
[PubMed]
14.
van der Straat
L.
,
Vernooij
M.
,
Lammers
M.
,
van den Berg
W.
,
Schonewille
T.
,
Cordewener
J.
et al.
(
2014
)
Expression of the Aspergillus terreus itaconic acid biosynthesis cluster in Aspergillus niger
.
Microb. Cell Fact
13
,
11
[PubMed]
15.
Cairns
T.C.
,
Nai
C.
and
Meyer
V.
(
2018
)
How a fungus shapes biotechnology: 100 years of Aspergillus niger research
.
Fungal Biol. Biotechnol.
5
,
1
14
[PubMed]
17.
Meyer
V.
(
2021
)
Metabolic Engineering of Filamentous Fungi
.
Metabolic Engineering: Concepts and Applications
1st Ed.
Wiley-VCH GmbH
.
In Press
18.
Boecker
S.
,
Grätz
S.
,
Kerwat
D.
,
Adam
L.
,
Schirmer
D.
,
Richter
L.
et al.
(
2018
)
Aspergillus niger is a superior expression host for the production of bioactive fungal cyclodepsipeptides
.
Fungal Biol. Biotechnol.
5
,
4
.
[PubMed]
19.
Richter
L.
,
Wanka
F.
,
Boecker
S.
,
Storm
D.
,
Kurt
T.
,
Vural
Ö.
et al.
(
2014
)
Engineering of Aspergillus niger for the production of secondary metabolites
.
Fungal Biol. Biotechnol.
1
,
4
,
[PubMed]
20.
German-Fattal
M.
(
2001
)
Fusafungine, an antimicrobial with anti-inflammatory properties in respiratory tract infections
.
Clin. Drug Invest.
21
,
653
670
21.
Steiniger
C.
,
Hoffmann
S.
,
Mainz
A.
,
Kaiser
M.
,
Voigt
K.
,
Meyer
V.
et al.
(
2017
)
Harnessing fungal nonribosomal cyclodepsipeptide synthetases for mechanistic insights and tailored engineering
.
Chem. Sci.
8
,
7834
7843
.
[PubMed]
22.
Cortesão
M.
,
de Haas
A.
,
Unterbusch
R.
,
Fujimori
A.
,
Schütze
T.
,
Meyer
V.
et al.
(
2020
)
Aspergillus niger spores are highly resistant to space radiation
.
Front Microbiol.
11
,
560
[PubMed]
23.
Checinska
A.
,
Probst
A.J.
,
Vaishampayan
P.
,
White
J.R.
,
Kumar
D.
,
Stepanov
V.G.
et al.
(
2015
)
Microbiomes of the dust particles collected from the International Space Station and Spacecraft assembly Facilities
.
Microbiome
3
,
50
[PubMed]
24.
Novikova
N.
,
De Boever
P.
,
Poddubko
S.
,
Deshevaya
E.
,
Polikarpov
N.
,
Rakova
N.
et al.
(
2006
)
Survey of environmental biocontamination on board the International Space Station
.
In: Res. Microbiol.
157
,
5
12
25.
Cortesão
M.
,
Schütze
T.
,
Marx
R.
,
Moeller
R.
and
Meyer
V.
(
2020
)
Fungal biotechnology in space: why and how?
In: Grand Challenges in Fungal Biotechnol.
501
535
26.
Romsdahl
J.
,
Blachowicz
A.
,
Chiang
Y.M.
,
Venkateswaran
K.
and
Wang
C.C.C.
(
2020
)
Metabolomic analysis of Aspergillus niger isolated from the International Space Station reveals enhanced production levels of the antioxidant pyranonigrin A
.
Front Microbiol.
11
,
931
[PubMed]
27.
Cairns
T.C.
,
Zheng
X.
,
Zheng
P.
,
Sun
J.
and
Meyer
V.
(
2019
)
Moulding the mould: understanding and reprogramming filamentous fungal growth and morphogenesis for next generation cell factories
.
Biotechnol. Biofuels
12
,
77
[PubMed]
28.
Upton
D.J.
,
McQueen-Mason
S.J.
and
Wood
A.J.
(
2017
)
An accurate description of Aspergillus niger organic acid batch fermentation through dynamic metabolic modelling
.
Biotechnol. Biofuels
10
,
[PubMed]
29.
Odoni
D.I.
,
van Gaal
M.P.
,
Schonewille
T.
,
Tamayo-Ramos
J.A.
,
dos Santos
V.A.P.
,
Suarez-Diez
M.
et al.
(
2017
)
Aspergillus niger secretes citrate to increase iron bioavailability
.
Front Microbiol.
8
,
1424
.
[PubMed]
30.
Witteveen
C.F.B.
,
Veenhuis
M.
and
Visser
J.
(
1992
)
Localization of glucose oxidase and catalase activities in Aspergillus niger
.
Appl. Environ. Microbiol.
58
,
1190
1194
[PubMed]
31.
Ruijter
G.J.G.
,
Panneman
H.
,
Xu
D.B.
and
Visser
J.
(
2000
)
Properties of Aspergillus niger citrate synthase and effects of citA overexpression on citric acid production
.
FEMS Microbiol. Lett.
184
,
35
40
,
1
[PubMed]
32.
Kirimura
K.
,
Kobayashi
K.
,
Ueda
Y.
and
Hattori
T.
(
2016
)
Phenotypes of gene disruptants in relation to a putative mitochondrial malate-citrate shuttle protein in citric acid-producing Aspergillus niger
.
Biosci. Biotechnol. Biochem.
90
,
1737
1746
,
9
33.
Steiger
M.G.
,
Rassinger
A.
,
Mattanovich
D.
and
Sauer
M.
(
2019
)
Engineering of the citrate exporter protein enables high citric acid production in Aspergillus niger
.
Metab. Eng.
52
,
224
231
[PubMed]
34.
Ruijter
G.J.G.
,
Panneman
H.
and
Visser
J.
(
1997
)
Overexpression of phosphofructokinase and pyruvate kinase in citric acid-producing Aspergillus niger
.
Biochim. Biophys. Acta
1334
,
317
326
[PubMed]
35.
Meyer
V.
,
Wanka
F.
,
van Gent
J.
,
Arentshorst
M.
,
van den Hondel
C.A.
and
Ram
A.F.
(
2011
)
Fungal gene expression on demand: an inducible, tunable, and metabolism-independent expression system for Aspergillus niger
.
Appl. Env. Microbiol.
77
,
2975
2983
.
36.
Wanka
F.
,
Cairns
T.
,
Boecker
S.
,
Berens
C.
,
Happel
A.
,
Zheng
X.
et al.
(
2016
)
Tet-on, or Tet-off, that is the question: Advanced conditional gene expression in Aspergillus
.
Fungal Genet. Biol.
89
,
72
83
[PubMed]
37.
Yang
L.
,
Linde
T.
,
Hossain
A.H.
,
Lübeck
M.
,
Punt
P.J.
and
Lübeck
P.S.
(
2019
)
Disruption of a putative mitochondrial oxaloacetate shuttle protein in Aspergillus carbonarius results in secretion of malic acid at the expense of citric acid production
.
BMC Biotechnol.
19
,
72
[PubMed]
38.
Yang
L.
,
Henriksen
M.M.
,
Hansen
R.S.
,
Lübeck
M.
,
Vang
J.
,
Andersen
J.E.
et al.
(
2020
)
Metabolic engineering of Aspergillus niger via ribonucleoprotein-based CRISPR–Cas9 system for succinic acid production from renewable biomass
.
Biotechnol. Biofuels
13
,
206
[PubMed]
39.
Wierckx
N.
,
Agrimi
G.
,
Lübeck
P.S.
,
Steiger
M.G.
,
Mira
N.P.
and
Punt
P.J.
(
2020
)
Metabolic specialization in itaconic acid production: a tale of two fungi
.
Curr. Opin. Biotechnol.
62
,
153
159
[PubMed]
40.
Hossain
A.H.
,
Li
A.
,
Brickwedde
A.
,
Wilms
L.
,
Caspers
M.
,
Overkamp
K.
et al.
(
2016
)
Rewiring a secondary metabolite pathway towards itaconic acid production in Aspergillus niger
.
Microb. Cell Fact
15
,
130
[PubMed]
41.
Hossain
A.H.
,
Ter Beek
A.
and
Punt
P.J.
(
2019
)
Itaconic acid degradation in Aspergillus niger: The role of unexpected bioconversion pathways
.
Fungal Biol. Biotechnol.
6
,
1
42.
Takeshita
N.
and
Fischer
R.
(
2011
)
On the role of microtubules, cell end markers, and septal microtubule organizing centres on site selection for polar growth in Aspergillus nidulans
.
Fungal Biol.
115
,
506
517
[PubMed]
43.
Takeshita
N.
,
Mania
D.
,
Herrero
S.
,
Ishitsuka
Y.
,
Nienhaus
G.U.
,
Podolski
M.
et al.
(
2013
)
The cell-end marker TeaA and the microtubule polymerase AlpA contribute to microtubule guidance at the hyphal tip cortex of Aspergillus nidulans to provide polarity maintenance
.
J. Cell Sci.
126
,
5400
5411
[PubMed]
44.
Ishitsuka
Y.
,
Savage
N.
,
Li
Y.
,
Bergs
A.
,
Grün
N.
,
Kohler
D.
et al.
(
2015
)
Superresolution microscopy reveals a dynamic picture of cell polarity maintenance during directional growth
.
Sci Adv.
1
,
e1500947
[PubMed]
45.
Riquelme
M.
,
Bredeweg
E.L.
,
Callejas-Negrete
O.
,
Roberson
R.W.
,
Ludwig
S.
,
Beltrán-Aguilar
A.
et al.
(
2014
)
The Neurospora crassa exocyst complex tethers Spitzenkörper vesicles to the apical plasma membrane during polarized growth
.
Mol. Biol. Cell.
25
,
1312
1326
[PubMed]
46.
Fiedler
M.R.M.
,
Cairns
T.C.
,
Koch
O.
,
Kubisch
C.
and
Meyer
V.
(
2018
)
Conditional expression of the small GTPase ArfA impacts secretion, morphology, growth, and actin ring position in Aspergillus niger
.
Front Microbiol.
9
,
878
.
[PubMed]
47.
Fiedler
M.R.M.
,
Barthel
L.
,
Kubisch
C.
,
Nai
C.
and
Meyer
V.
(
2018
)
Construction of an improved Aspergillus niger platform for enhanced glucoamylase secretion
.
Microb. Cell Fact.
17
,
95
[PubMed]
48.
Gordon
C.L.
,
Khalaj
V.
,
Ram
A.F.J.
,
Archer
D.B.
,
Brookman
J.L.
,
Trinci
A.P.J.
et al.
(
2000
)
Glucoamylase::green fluorescent protein fusions to monitor protein secretion in Aspergillus niger
.
Microbiology
146
,
415
426
[PubMed]
49.
Kwon
M.J.
,
Arentshorst
M.
,
Roos
E.D.
,
Van Den Hondel
C.A.
,
Meyer
V.
and
Ram
A.F.
(
2011
)
Functional characterization of Rho GTPases in Aspergillus niger uncovers conserved and diverged roles of Rho proteins within filamentous fungi
.
Mol. Microbiol.
79
,
1151
1167
[PubMed]
50.
Read
N.D.
(
2011
)
Exocytosis and growth do not occur only at hyphal tips
.
Mol. Microbiol.
81
,
4
7
[PubMed]
51.
Hayakawa
Y.
,
Ishikawa
E.
,
Shoji
J.
,
Nakano
H.
and
Kitamoto
K.
(
2011
)
Septum-directed secretion in the filamentous fungus Aspergillus oryzae
.
Mol. Microbiol.
81
,
40
55
[PubMed]
52.
Benoit
I.
,
Zhou
M.
,
Vivas Duarte
A.
,
Downes
D.J.
,
Todd
R.B.
,
Kloezen
W.
et al.
(
2015
)
Spatial differentiation of gene expression in Aspergillus niger colony grown for sugar beet pulp utilization
.
Sci. Rep.
5
,
[PubMed]
53.
Braaksma
M.
,
Martens-Uzunova
E.S.
,
Punt
P.J.
and
Schaap
P.J.
(
2010
)
An inventory of the Aspergillus niger secretome by combining in silico predictions with shotgun proteomics data
.
BMC Genomics
11
,
548
[PubMed]
54.
Burggraaf
A.M.
,
Punt
P.J.
and
Ram
A.F.J.
(
2016
)
The unconventional secretion of PepN is independent of a functional autophagy machinery in the filamentous fungus Aspergillus niger
.
FEMS Microbiol. Lett.
363
,
[PubMed]
55.
Rodrigues
M.L.
,
Nakayasu
E.S.
,
Oliveira
D.L.
,
Nimrichter
L.
,
Nosanchuk
J.D.
,
Almeida
I.C.
et al.
(
2008
)
Extracellular vesicles produced by Cryptococcus neoformans contain protein components associated with virulence
.
Eukaryot Cell.
7
,
58
67
[PubMed]
56.
Gibson
R.K.
and
Peberdy
J.F.
(
1972
)
Fine structure of protoplasts of Aspergillus nidulans
.
J. Gen. Microbiol.
72
,
3
57.
Bleackley
M.R.
,
Dawson
C.S.
and
Anderson
M.A.
(
2019
)
Fungal extracellular vesicles with a focus on proteomic analysis
.
Proteomics
19
,
8
58.
Wanka
F.
,
Arentshorst
M.
,
Cairns
T.C.
,
Jørgensen
T.
,
Ram
A.F.J.
and
Meyer
V.
(
2016
)
Highly active promoters and native secretion signals for protein production during extremely low growth rates in Aspergillus niger
.
Microb. Cell Fact.
15
,
145
,
[PubMed]
59.
Keller
N.P.
,
Turner
G.
and
Bennett
J.W.
(
2005
)
Fungal secondary metabolism: from biochemistry to genomics
.
Nat. Rev. Micro.
3
,
937
947
60.
Pel
H.J.
,
De Winde
J.H.
,
Archer
D.B.
,
Dyer
P.S.
,
Hofmann
G.
,
Schaap
P.J.
et al.
(
2007
)
Genome sequencing and analysis of the versatile cell factory Aspergillus niger CBS 513.88
.
Nat. Biotechnol.
25
,
221
231
[PubMed]
61.
Inglis
D.O.
,
Binkley
J.
,
Skrzypek
M.S.
,
Arnaud
M.B.
,
Cerqueira
G.C.
,
Shah
P.
et al.
(
2013
)
Comprehensive annotation of secondary metabolite biosynthetic genes and gene clusters of Aspergillus nidulans, A. fumigatus, A. niger and A. oryzae
.
BMC Microbiol.
13
,
91
.
[PubMed]
62.
Schäpe
P.
,
Kwon
M.J.
,
Baumann
B.
,
Gutschmann
B.
,
Jung
S.
,
Lenz
S.
et al.
(
2019
)
Updating genome annotation for the microbial cell factory Aspergillus niger using gene co-expression networks
.
Nucleic Acids Res.
47
,
2
63.
Nai
C.
and
Meyer
V.
(
2017
)
From axenic to mixed cultures: Technological advances accelerating a paradigm shift in microbiology
.
Trends Microbiol.
26
,
538
554
[PubMed]
64.
Macheleidt
J.
,
Mattern
D.J.
,
Fischer
J.
,
Netzker
T.
,
Weber
J.
,
Schroeckh
V.
et al.
(
2016
)
Regulation and role of fungal secondary metabolites
.
Annu. Rev. Genet.
50
,
371
392
[PubMed]
65.
Martín
J.F.
(
2020
)
Transport systems, intracellular traffic of intermediates and secretion of β-lactam antibiotics in fungi
.
Fungal Biol. Biotechnol.
7
,
6
[PubMed]
66.
Herr
A.
and
Fischer
R.
(
2014
)
Improvement of Aspergillus nidulans penicillin production by targeting AcvA to peroxisomes
.
Metab. Eng.
25
,
131
139
[PubMed]
67.
Sui
Y.F.
,
Schütze
T.
,
Ouyang
L.M.
,
Lu
H.
,
Liu
P.
,
Xiao
X.
et al.
(
2020
)
Engineering cofactor metabolism for improved protein and glucoamylase production in Aspergillus niger
.
Microb. Cell Fact.
19
,
198
[PubMed]
68.
Niu
J.
,
Arentshorst
M.
,
Nair
P.D.S.
,
Dai
Z.
,
Baker
S.E.
,
Frisvad
J.C.
et al.
(
2016
)
Identification of a classical mutant in the industrial host Aspergillus niger by systems genetics: LaeA is required for citric acid production and regulates the formation of some secondary metabolites
.
G3 Genes|Genomes|Genetics
6
,
193
204
.
69.
Van Den Hombergh
J.P.T.W.
,
Sollewijn Gelpke
M.D.
,
Van De Vondervoort
P.J.I.
,
Buxton
F.P.
and
Visser
J.
(
1997
)
Disruption of three acid proteases in Aspergillus niger: Effects on protease spectrum, intracellular proteolysis, and degradation of target proteins
.
Eur. J. Biochem.
247
,
2
70.
Arentshorst
M.
and
Ram
A.F.J.
(
2018
)
Parasexual crossings for bulk segregant analysis in Aspergillus niger to facilitate mutant identification via whole genome sequencing
.
In: Methods in Mol. Biol.
1775
,
277
287
71.
Kadooka
C.
,
Nakamura
E.
,
Mori
K.
,
Okutsu
K.
,
Yoshizaki
Y.
,
Takamine
K.
et al.
(
2020
)
LaeA controls citric acid production through regulation of the citrate exporter-encoding gene in Aspergillus luchuensis
.
Appl. Environ. Microbiol.
86
,
e01950
e02019
,
[PubMed]
72.
Wang
B.
,
Lv
Y.
,
Li
X.
,
Lin
Y.
,
Deng
H.
and
Pan
L.
(
2018
)
Profiling of secondary metabolite gene clusters regulated by LaeA in Aspergillus niger FGSC A1279 based on genome sequencing and transcriptome analysis
.
Res. Microbiol.
169
,
67
77
[PubMed]
73.
Pirt
S.J.
(
1966
)
A Theory of the mode of growth of fungi in the form of pllets in submerged culture
.
Proc. R. Soc. Lond. B.
166
,
369
373
[PubMed]
74.
Veiter
L.
,
Rajamanickam
V.
and
Herwig
C.
(
2018
)
The filamentous fungal pellet—relationship between morphology and productivity
.
Appl. Microbiol. Biotechnol.
102
,
2997
3006
[PubMed]
75.
Cairns
T.C.
,
Feurstein
C.
,
Zheng
X.
,
Zheng
P.
,
Sun
J.
and
Meyer
V.
(
2019
)
A quantitative image analysis pipeline for the characterization of filamentous fungal morphologies as a tool to uncover targets for morphology engineering: a case study using aplD in Aspergillus niger
.
Biotechnol. Biofuels
12
,
149
[PubMed]
76.
Cairns
T.C.
,
Feurstein
C.
,
Zheng
X.
,
Zhang
L.H.
,
Zheng
P.
,
Sun
J.
et al.
(
2019
)
Functional exploration of co-expression networks identifies a nexus for modulating protein and citric acid titres in Aspergillus niger submerged culture
.
Fungal Biol. Biotechnol.
6
,
18
[PubMed]
77.
Brakhage
A.A.
(
2013
)
Regulation of fungal secondary metabolism
.
Nat. Rev. Micro.
11
,
21
32
78.
Papagianni
M.
(
2004
)
Fungal morphology and metabolite production in submerged mycelial processes
.
Biotechnol. Adv.
22
,
3
79.
Lin
P.J.
,
Scholz
A.
and
Krull
R.
(
2010
)
Effect of volumetric power input by aeration and agitation on pellet morphology and product formation of Aspergillus niger
.
Biochem. Eng. J.
49
,
2
80.
Hille
A.
,
Neu
T.R.
,
Hempel
D.C.
and
Horn
H.
(
2005
)
Oxygen profiles and biomass distribution in biopellets of Aspergillus niger
.
Biotechnol. Bioeng.
5
,
614
623
,
92
81.
Villena
G.K.
and
Gutiérrez-Correa
M.
(
2007
)
Morphological patterns of Aspergillus niger biofilms and pellets related to lignocellulolytic enzyme productivities
.
Lett. Appl. Microbiol.
45
,
231
237
[PubMed]
82.
Ehgartner
D.
,
Herwig
C.
and
Fricke
J.
(
2017
)
Morphological analysis of the filamentous fungus Penicillium chrysogenum using flow cytometry—the fast alternative to microscopic image analysis
.
Appl. Microbiol. Biotechnol.
101
,
7675
7688
[PubMed]
83.
Schmideder
S.
,
Barthel
L.
,
Friedrich
T.
,
Thalhammer
M.
,
Kovačević
T.
,
Niessen
L.
et al.
(
2019
)
An X-ray microtomography-based method for detailed analysis of the three-dimensional morphology of fungal pellets
.
Biotechnol. Bioeng.
116
,
6
84.
Schmideder
S.
,
Barthel
L.
,
Müller
H.
,
Meyer
V.
and
Briesen
H.
(
2019
)
From three-dimensional morphology to effective diffusivity in filamentous fungal pellets
.
Biotechnol. Bioeng.
116
,
12
85.
Schmideder
S.
,
Müller
H.
,
Barthel
L.
,
Friedrich
T.
,
Niessen
L.
,
Meyer
V.
et al.
(
2020
)
Universal law for diffusive mass transport through mycelial networks
.
Biotechnol. Bioeng.
118
,
930
943
[PubMed]
86.
Meyer
V.
,
Arentshorst
M.
,
El-Ghezal
A.
,
Drews
A.C.
,
Kooistra
R.
,
van den Hondel
C.A.
et al.
(
2007
)
Highly efficient gene targeting in the Aspergillus niger kusA mutant
.
J. Biotechnol.
128
,
770
775
.
[PubMed]
87.
De Sena-Tomas
C.
,
Yu
E.Y.
,
Calzada
A.
,
Holloman
W.K.
,
Lue
N.F.
and
Perez-Martin
J.
(
2015
)
Fungal Ku prevents permanent cell cycle arrest by suppressing DNA damage signaling at telomeres
.
Nucleic Acids Res.
43
,
2138
2151
[PubMed]
88.
Vanegas
K.G.
,
Jarczynska
Z.D.
,
Strucko
T.
and
Mortensen
U.H.
(
2019
)
Cpf1 enables fast and efficient genome editing in Aspergilli
.
Fungal Biol. Biotechnol.
6
,
1
10
89.
Zheng
X.
,
Zheng
P.
,
Zhang
K.
,
Cairns
T.C.
,
Meyer
V.
,
Sun
J.
et al.
(
2018
)
5S rRNA Promoter for guide RNA expression enabled highly efficient CRISPR/Cas9 genome editing in Aspergillus niger
.
ACS Synth. Biol.
8
,
1568
1574
[PubMed]
90.
Kwon
M.J.
,
Schütze
T.
,
Spohner
S.
,
Haefner
S.
and
Meyer
V.
(
2019
)
Practical guidance for the implementation of the CRISPR genome editing tool in filamentous fungi
.
Fungal Biol. Biotechnol.
6
,
15
[PubMed]
91.
Arunmozhi Balajee
S.
,
Imhof
A.
,
Gribskov
J.L.
and
Marr
K.A.
(
2005
)
Determination of antifungal drug susceptibilities of Aspergillus species by a fluorenscence-based microplate assay
.
J. Antimicrob. Chemother.
55
,
102
105
[PubMed]
92.
Alberto
F.
,
Navarro
D.
,
De Vries
R.P.
,
Asther
M.
and
Record
E.
(
2009
)
Technical advance in fungal biotechnology: Development of a miniaturized culture method and an automated high-throughput screening
.
Lett. Appl. Microbiol.
49
,
278
282
[PubMed]
93.
Hevekerl
A.
,
Kuenz
A.
and
Vorlop
K.D.
(
2014
)
Filamentous fungi in microtiter plates - An easy way to optimize itaconic acid production with Aspergillus terreus
.
Appl. Microbiol. Biotechnol.
98
,
6983
6989
[PubMed]
94.
Zhang
L.
,
Zheng
X.
,
Cairns
T.C.
,
Zhang
Z.
,
Wang
D.
,
Zheng
P.
et al.
(
2020
)
Disruption or reduced expression of the orotidine-5′-decarboxylase gene pyrG increases citric acid production: a new discovery during recyclable genome editing in Aspergillus niger
.
Microb. Cell Fact.
19
,
76
[PubMed]
95.
Huth
I.
,
Schrader
J.
and
Holtmann
D.
(
2017
)
Microtiter plate-based cultivation to investigate the growth of filamentous fungi
.
Eng. Life Sci.
17
,
10
96.
Jansen
R.P.
,
Beuck
C.
,
Moch
M.
,
Klein
B.
,
Küsters
K.
,
Morschett
H.
et al.
(
2019
)
A closer look at Aspergillus: online monitoring via scattered light enables reproducible phenotyping
.
Fungal Biol. Biotechnol.
6
,
11
[PubMed]
97.
Jansen
R.
,
Küsters
K.
,
Morschett
H.
,
Wiechert
W.
and
Oldiges
M.
(
2021
)
A fully automated pipeline for the dynamic at‐line morphology analysis of microscale Aspergillus cultivation
.
Fungal Biol. Biotechnol.
8
,
2
[PubMed]
98.
Carvalho
N.D.S.P.
,
Arentshorst
M.
,
Jin Kwon
M.
,
Meyer
V.
and
Ram
A.F.J.
(
2010
)
Expanding the ku70 toolbox for filamentous fungi: establishment of complementation vectors and recipient strains for advanced gene analyses
.
Appl. Microbiol. Biotechnol.
87
,
1463
1473
[PubMed]
99.
Andersen
M.R.
,
Salazar
M.P.
,
Schaap
P.J.
,
Van De Vondervoort
P.J.I.
,
Culley
D.
et al.
(
2011
)
Comparative genomics of citric-acid-producing Aspergillus niger ATCC 1015 versus enzyme-producing CBS 513.88
.
Genome Res.
21
,
885
897
[PubMed]
100.
Yin
C.
,
Wang
B.
,
He
P.
,
Lin
Y.
and
Pan
L.
(
2014
)
Genomic analysis of the aconidial and high-performance protein producer, industrially relevant Aspergillus niger SH2 strain
.
Gene
541
,
107
114
[PubMed]
101.
Gong
W.
,
Cheng
Z.
,
Zhang
H.
,
Liu
L.
,
Gao
P.
and
Wang
L.
(
2016
)
Draft Genome Sequence of Aspergillus niger Strain An76
.
Genome Announc.
4
,
e01700
01715
[PubMed]
102.
Gong
W.
,
Dai
L.
,
Zhang
H.
,
Zhang
L.
and
Wang
L.
(
2018
)
A Highly efficient xylan-utilization system in Aspergillus niger An76: A functional-proteomics study
.
Front Microbiol.
9
,
430
[PubMed]
103.
Paul
S.
,
Ludeña
Y.
,
Villena
G.K.
,
Yu
F.
,
Sherman
D.H.
and
Gutiérrez-Correa
M.
(
2017
)
High-quality draft genome sequence of a biofilm forming lignocellulolytic Aspergillus niger strain ATCC 10864
.
Stand Genomic Sci.
12
,
104.
Sui
Y.F.
,
Ouyang
L.M.
,
Schütze
T.
,
Cheng
S.
,
Meyer
V.
and
Zhuang
Y.P.
(
2020
)
Comparative genomics of the aconidial Aspergillus niger strain LDM3 predicts genes associated with its high protein secretion capacity
.
Appl. Microbiol. Biotechnol.
104
,
2623
2637
[PubMed]
This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY).