Next Article in Journal
Phylogenetic Review of Acaulospora (Diversisporales, Glomeromycota) and the Homoplasic Nature of Its Ornamentations
Next Article in Special Issue
Investigating the Effect of Alcohol Dehydrogenase Gene Knockout on Lipid Accumulation in Mucor circinelloides WJ11
Previous Article in Journal
Characterisation of Pythium aristosporum Oomycete—A Novel Pathogen Causing Rice Seedling Blight in China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Comparative Genomics of Mortierellaceae Provides Insights into Lipid Metabolism: Two Novel Types of Fatty Acid Synthase

1
College of Life Sciences, Shandong Normal University, Jinan 250358, China
2
Institute of Microbiology, School of Ecology and Nature Conservation, Beijing Forestry University, Beijing 100083, China
3
School of Civil Engineering and Architecture, Anhui University of Technology, Ma’anshan 243002, China
4
State Key Laboratory of Mycology, Institute of Microbiology, Chinese Academy of Sciences, Beijing 100101, China
*
Author to whom correspondence should be addressed.
J. Fungi 2022, 8(9), 891; https://doi.org/10.3390/jof8090891
Submission received: 24 July 2022 / Revised: 16 August 2022 / Accepted: 22 August 2022 / Published: 23 August 2022
(This article belongs to the Special Issue New Perspectives for Oleaginous Fungi)

Abstract

:
Fungal species in the family Mortierellaceae are important for their remarkable capability to synthesize large amounts of polyunsaturated fatty acids, especially arachidonic acid (ARA). Although many genomes have been published, the quality of these data is not satisfactory, resulting in an incomplete understanding of the lipid pathway in Mortierellaceae. We provide herein two novel and high-quality genomes with 55.32% of syntenic gene pairs for Mortierella alpina CGMCC 20262 and M. schmuckeri CGMCC 20261, spanning 28 scaffolds of 40.22 Mb and 25 scaffolds of 49.24 Mb, respectively. The relative smaller genome for the former is due to fewer protein-coding gene models (11,761 vs. 13,051). The former yields 45.57% of ARA in total fatty acids, while the latter 6.95%. The accumulation of ARA is speculated to be associated with delta-5 desaturase (Delta5) and elongation of very long chain fatty acids protein 3 (ELOVL3). A further genomic comparison of 19 strains in 10 species in three genera in the Mortierellaceae reveals three types of fatty acid synthase (FAS), two of which are new to science. The most common type I exists in 16 strains of eight species of three genera, and was discovered previously and consists of a single unit with eight active sites. The newly revealed type II exists only in M. antarctica KOD 1030 where the unit is separated into two subunits α and β comprised of three and five active sites, respectively. Another newly revealed type III exists in M. alpina AD071 and Dissophora globulifera REB-010B, similar to type II but different in having one more acyl carrier protein domain in the α subunit. This study provides novel insights into the enzymes related to the lipid metabolism, especially the ARA-related Delta5, ELOVL3, and FAS, laying a foundation for genetic engineering of Mortierellaceae to modulate yield in polyunsaturated fatty acids.

1. Introduction

Environmental protection and sustainable development are now giving urgency to identifying alternatives for oil plants and animals. Therefore, oleaginous fungi such as Aspergillus flavus Link [=A. oryzae (Ahlb.) Cohn], Mortierella alpina Peyronel, M. wolfii B.S. Mehrotra & Baijal, Mucor racemosus Fresen., Lichtheimia corymbifera (Cohn) Vuill., Syncephalastrum racemosum Cohn ex J. Schröt., Geotrichum candidum Link, Umbelopsis isabellina (Oudem.) W. Gams (=Mortierella isabellina Oudem.), and Yarrowia lipolytica (Wick., Kurtzman & Herman) Van der Walt & Arx have recently gained attention and played an important role in the studies of biodiesel and polyunsaturated fatty acids (PUFAs) [1,2,3,4,5,6,7,8,9,10,11]. These oleaginous fungi mainly belong to Ascomycota, Mortierellomycota, and Mucoromycota.
Compared with other microorganisms such as microalgae and algae, oleaginous fungi are distinct in possessing a shorter life cycle, assimilating a variety of carbon sources, and adapting to various weather and seasons [1,7,9]. Meanwhile, the species in the Mortierellaceae of Mortierellomycota are characterized by synthesizing a large amount of PUFAs [1,12]. At present, 122 species of seven genera are accommodated in the family Mortierellaceae [13], and almost all tested strains of Mortierellaceae were found to be able to synthesize arachidonic acid (ARA, C20:4) [1,12,14,15,16], which is a vital nutrient for the elderly and infants, possessing multiple functions such as the protection of brain and muscles and potential against tumors and inflammation [16,17,18].
Mortierella alpina synthesizing lipids more than 50% of its dry cell weight and ARA accounting for 30–70% of the total fatty acids has recently been one of the important model organisms for PUFA metabolism [12,16,19,20,21,22]. Based on genetic manipulation in M. alpina, numerous enzymes such as AMP deaminase, omega-3 desaturase, delta-5 desaturase, and fatty acid synthase were proved to play key functions in the biosynthesis of PUFAs [12,16,23,24,25]. For example, overexpression of homologous AMP deaminase in M. alpina resulted in a significant increase of 15.0–34.3% in lipid yield [23].
In 2011, the first draft genome of Mortierella alpina was completed for the strain ATCC 32222, initializing studies on lipid metabolic pathway in the species [11]. Until now, nine genomes of M. alpina have been published in the NCBI database (https://www.ncbi.nlm.nih.gov, accessed on 2 July 2022) [12,26,27,28]. Other genomes of the Mortierellaceae were assembled for Actinomortierella wolfii (B.S. Mehrotra & Baijal) Vandepol & Bonito, Dissophora globulifera (O. Rostr.) Vandepol & Bonito, Mortierella amoeboidea W. Gams, M. antarctica Linnem., M. elongata Linnem., M. epicladia W. Gams & Emden, M. gamsii Milko, and M. verticillata Linnem. [27,29,30,31]. However, comprehensive analyses of these genomes, especially the key genes involved in fatty acid synthesis, are not satisfactory due to a lack of good-quality sequence data.
In the present study, we sequence two strains, Mortierella alpina CGMCC 20262 and M. schmuckeri Linnem. CGMCC 20261, by combining next-generation with PacBio sequencing. A total of 40 strains of basal fungi are used to reconstruct their phylogenomic relationship, and 13 protein-coding genes involved in the lipid metabolism are characterized in 19 strains of Mortierellaceae.

2. Materials and Methods

2.1. Strains, Media, and Fermentation

Mortierella alpina CGMCC 20262 and M. schmuckeri CGMCC 20261 were deposited in the China General Microbiological Culture Collection Center, Beijing, China (CGMCC). Cultures were incubated with potato dextrose agar (PDA: 200 g/L potato, 20 g/L glucose, 20 g/L agar, and 1000 mL distilled water) and ampicillin (100 μg/mL) at 20 °C for ten days. A shake flask fermentation was then carried out following a previous study [1]. In brief, 1 mL of spore suspension (1 × 106) was incubated at 20 °C and 140 rpm for seven days in a 250 mL flask with 100 mL of modified Kendrick media (50 g/L glucose, 2 g/L diammonium tartrate, 7 g/L KH2PO4, 2 g/L Na2HPO4, 1.5 g/L MgSO4·7H2O, 1.5 g/L yeast extract, 0.1 g/L CaCl2·2H2O, 8 mg/L FeCl3·6H2O, 1 mg/L ZnSO4·7H2O, 0.1 mg/L CuSO4·5H2O, 0.1 mg/L CO(NO3)2·6H2O, 0.1 mg/L MnSO4·5H2O, and pH 6.0) [1,23,32].

2.2. Biomass, Fatty Acid Measurement, and Profiling

Biomass, fatty acid measurement, and profiling followed the method by Zhao et al. [1]. Fresh biomasses were drained with a vacuum pump, and then freeze-dried for two days to attain a constant dry cell weight (DCW). The freeze-dried biomasses were hydrolyzed with HCl solution (6 mol/L), and then extracted with ethanol, anhydrous ether, and petroleum ether following previous studies [1,33]. The extract was then dried with an oven by slowly warming to 80 °C until a constant total lipid weight (TLW) was attained. Total lipid content (TLC) was calculated as the TLW being divided by the DCW. For fatty acid profiling, the gas chromatography–mass spectrometry (GC/MS, QP2010, Shimadzu Corp., Japan) was performed with a 30 m × 0.25 mm × 0.25 μm column (Rtx-5MS, RESTEK, Bellefonte, PA, USA), and nonanoic acid (C9:0) was selected as internal standard according to Zhao et al. [1].

2.3. Genome Sequencing and Assembly

Mortierella alpina CGMCC 20262 and M. schmuckeri CGMCC 20261 were incubated with PDA at 20 °C for five days. Total cell DNAs were extracted from mycelia using a kit (O-GPLF-400, GeneOnBio Corporation, Changchun, China) according to the manufacturer’s protocol, and then detected by DNA/Protein Analyzer and 1% agarose gel electrophoresis. High-quality DNAs were sequenced at Beijing Novogene Bioinformatics Technology Co., Ltd. (Beijing, China) by using a PacBio Sequel and Illumina NovaSeq 6000 platform with 20 kb and 350 bp library, respectively. Low-quality reads (less than 500 bp) were removed by quality control from the raw data produced with the PacBio Sequel platform, and then the controlled high-quality reads were de novo assembled using SMRT Link v5.1.0 [34]. The assembled genomes were assessed by Quast v5.0.2 [35] and BUSCO v5.2.2 [36].

2.4. Gene Prediction and Functional Annotation

Protein-coding gene models of Mortierella alpina CGMCC 20262 and M. schmuckeri CGMCC 20261 were de novo predicted using Augustus v3.3.3 [37]. Amino acid and DNA sequences were functionally annotated using NR (https://www.ncbi.nlm.nih.gov/protein/, accessed on 2 July 2022), NT (https://www.ncbi.nlm.nih.gov/nucleotide/, accessed on 2 July 2022), Pfam [38], GO [39], KEGG [40], CAZymes [41,42], and Eggnog [43] databases. All sequences were mapped onto these databases using Diamond v2.0.1 [44] with an e-value less than 1 × 10−5. The antiSMASH fungal version [45] was used to annotate the gene clusters of secondary metabolites with default parameters. Repetitive elements were identified using the Extensive de novo TE Annotator (EDTA) pipeline v1.9.5 [46]. RNAmmer v1.2 [47] and tRNAscan-SE v2.0.5 [48] were used to predict rRNAs and tRNAs, respectively.

2.5. Phylogenomic and Phylogenetic Analyses

A total of 40 strains of early diverging fungi, including 38 strains downloaded from online databases and two strains sequenced herein, were used for phylogenomic analyses (Table 1). A total of 192 clusters of orthologous proteins were identified with HMMER v3.3.1 [49] and Trimal v1.4.4 [50], following the methods described by Spatafora et al. [51] and James et al. [52], and then their amino acid sequences were aligned with MAFFT v7 [53]. Phylogenomic analyses were carried out with a Maximum Likelihood (ML) algorithm using RaxML v8.1.12 [54]. Maximum Likelihood analyses adopted the PROTGAMMALGX substitution model with 100 bootstrap replications. Phylogenetic analyses on amino acids of delta-5 desaturase and fatty acid synthase were performed using IQ-TREE v1.0 [55] with an ML algorithm, WAG substitution model, and 1000 bootstrap replications.

2.6. Comparative Genomic Analyses

The genome sequences of Mortierella alpina CGMCC 20262 and M. schmuckeri CGMCC 20261 were aligned using MCScanX for all protein-coding gene models, and then genomic collinearity was analyzed using a dual synteny plotter package [61]. For comparison of genes in lipid metabolism, especially those encoding fatty acid synthase and delta-5 desaturase, all 19 genomes of Mortierellaceae were reannotated using the KEGG database [41].

3. Results

3.1. Fatty Acid Profiles of Mortierella alpina and M. schmuckeri

Mortierella alpina CGMCC 20262 is much lower than M. schmuckeri CGMCC 20261 in dry cell weight (DCW, 6.6 g/L vs. 11.7 g/L), total lipid weight (TLW, 1.1 g/L vs. 8.0 g/L), and total lipid content (TLC, 17% vs. 70%; Figure 1a). However, M. alpina CGMCC 20262 yields 75.54% of polyunsaturated fatty acids (PUFAs; 15.41% of C18:2, 8.78% of C18:3, 3.40% of C20:3, 45.57% of C20:4, and 2.39% of C20:5) in total fatty acids (Figure 1b), more than the percentage of 23.03% in M. schmuckeri CGMCC 20261 (8.30% C18:2, 4.36% of C18:3, 3.43% of C20:3, 6.95% of C20:4, and no C20:5). Obviously, these two strains are significantly different, C20:4 (45.57%) being the top one in M. alpina CGMCC 20262, and C18:1 (27.54%) and C16:0 (24.68%) being the top two in M. schmuckeri CGMCC 20261 (Figure 1b).

3.2. Genomic Features of Mortierella alpina and M. schmuckeri

Assembled genomes span 28 scaffolds of 40.22 Mb with GC content of 50.92% in Mortierella alpina CGMCC 20262, and 25 scaffolds of 49.24 Mb with GC content of 47.46% in M. schmuckeri CGMCC 20261 (Table 2). The number of predicted protein-coding gene models of M. schmuckeri CGMCC 20261 is slightly more than that of M. alpina CGMCC 20262 (13,051 vs. 11,761; Table 2). Among these gene models, 66.17%, 33.97%, 47.35%, 42.69%, 73.20%, 37.71%, and 1.70% are mapped onto Pfam, NT, NR, GO, Eggnog, KEGG, and CAZymes databases in M. schmuckeri CGMCC 20261, and 72.33%, 34.23%, 41.70%, 46.01%, 77.17%, 41.31%, and 1.95% in M. alpina CGMCC 20262 (Table 2). Repetitive elements of M. alpina CGMCC 20262 and M. schmuckeri CGMCC 20261 account for 7.29% and 7.45% of the whole genomes (Table 2). Furthermore, 38 rRNAs (11 of 8s rRNA, 12 of 18s rRNA, and 15 of 28s rRNA) and 29 rRNAs (9 of 8s rRNA, 8 of 18s rRNA, and 12 of 28s rRNA) are predicted in M. alpina CGMCC 20262 and M. schmuckeri CGMCC 20261, respectively. More characteristics of genomes are listed in Table 2.

3.3. Phylogenomic Placements of Mortierella alpina and M. schmuckeri

The Maximum Likelihood phylogenomic tree (Figure 2) suggests that Mortierella alpina CGMCC 20262 is closely related to other strains of M. alpina (Maximum Likelihood bootstrap values, MLBV = 100%), and M. schmuckeri CGMCC 20261 is a sister to M. gamsii (MLBV = 100%) and closely related to M. elongata (MLBV = 100%).

3.4. Synteny between Mortierella alpina and M. schmuckeri

A total of 13,726 gene models located in all scaffolds are collinear between Mortierella alpina CGMCC 20262 and M. schmuckeri CGMCC 20261, accounting for 55.32% of all the 24,812 gene models in both strains. Figure 3 shows the synteny of the ten largest scaffolds where 16,927 gene models are predicted and 11,777 (69.58%) gene models are syntenic.

3.5. Lipid Metabolism in Mortierellaceae

To explore the lipid metabolism in Mortierellaceae, genomes of 19 strains in 10 species are de novo annotated or reannotated. We focus on the 13 genes associated with NADPH, precursors, desaturation, and elongation (Figure 4 and Supplementary Figure S2). The gene Delta5 (encoding delta-5 desaturase) is not found in M. gamsii NVP60, and neither is Delta6 (encoding delta-6 desaturase) in M. alpina LL118. With these two exceptions, each gene possesses between one and six copies. The genes HK (encoding hexokinase, a key enzyme in EMP or the glycolytic pathway) and ME (encoding malate dehydrogenase) are remarkable as both have between three and six copies (Supplementary Figure S2 and Supplementary Table S1).
Three types of fatty acid synthase are identified in the family Mortierellaceae (Figure 5, Supplementary File S1). Type I is ubiquitous (Actinomortierella wolfii NRRL 6351, Mortierella alpina AD072, M. alpina ATCC 32222, M. alpina B6842, M. alpina CCTCC M-207067, M. alpina CGMCC 20262, M. alpina CK1249, M. alpina GBA31, M. alpina LL118, M. alpina NRRL 66262, M. amoeboidea CBS 889.72, M. elongata AG-77, M. epicladia AD058, M. schmuckeri CGMCC 20261, and M. verticillata NRRL 6337), and unites together eight active sites (Figure 5). Type II is found in M. antarctica KOD 1030 only and includes two subunits, α subunit involving KR, KS, and PPT active sites, and β subunit comprising AT, ER, DH, MPT, and ACP active sites. Type III is identified in Dissophora globulifera REB-010B and Mortierella alpina AD071, similar to type II but with one more ACP active site in α subunit.
Four desaturases, i.e., Delta5, Delta6, Delta9, and Delta12, are found in the 19 strains of Mortierellaceae (Supplementary Figure S2 and Supplementary Table S1). Among these, the Delta5 catalyzes the synthesis of arachidonic acid (ARA, C20:4; Figure 4), which has diverged into three clades in Mortierellaceae (Figure 6 and Supplementary File S2).
The elongation of very long chain fatty acids protein (ELOVL) family consists of four members in Mortierellaceae, i.e., ELOVL2, ELOVL3, ELOVL4, and ELOVL6 (Supplementary Figure S1 and Supplementary File S3). ELOVL2 is shared by all strains of Mortierellaceae, while ELOVL3, ELOVL4, and ELOVL6 are species-specific. In detail, ELOVL3 is identified in M. alpina and M. amoeboidea, and ELOVL4 and ELOVL6 are annotated in M. verticillata and D. globulifera.

4. Discussion

Among numerous oleaginous fungi, Mortierella alpina is now attracting more and more attention because of its ability to accumulate large amounts of arachidonic acid (ARA) [12,16,22]. Previous studies suggested that M. alpina synthesized lipids up to 50% of dry cell weight (DCW) and ARA 30–70% of total lipid content (TLC) [12]. Herein, M. alpina CGMCC 20262 falls in these ranges, yielding lipids 17% of DCW and ARA 45.57% of TCL (Figure 1). In addition, we find a novel oleaginous fungus, M. schmuckeri CGMCC 20261, which accumulates lipids 70% of DCW (Figure 1), much more than M. alpina (17%).
The draft genome of M. alpina was published in 2011, and consequently lipid synthesis pathway was predicted [12]. Since then, more genomes of members in the family Mortierellaceae have shed light on their lipid synthesis mechanism. For example, Mortierella sp. BCC40632 has an n-6 series fatty acid synthesis pathway due to a lack of delta-15 or omega-3 desaturase [25,26,27,30,31]. In this paper, the genomes of two more strains, M. alpina CGMCC 20262 and M. schmuckeri CGMCC 20261, are sequenced to provide more insights into lipid metabolism (Table 2).
A large number of genes hitherto have been found to affect lipid synthesis in Mortierella alpina, such as acetyl-CoA carboxylase, AMP deaminase, malate dehydrogenase, delta-5 desaturase, delta-6 desaturase, and elongase 2 [12,22,23,66,67,68,69]. In this study, we compare 13 genes of 19 strains in Mortierellaceae (Supplementary Figure S2), finding differences in genes encoding hexokinase, delta-5 desaturase, and elongation of very long chain fatty acids protein. Compared with M. schmuckeri CGMCC 20261, the M. alpina CGMCC 20262 possesses one more copy of the HK gene, more copies of Delta5 and ELOVL genes but fewer copies of the Delta9 gene (Supplementary Figure S2 and Supplementary Table S1). Delta-5 desaturase and elongation of very long chain fatty acids protein play an essential role in the synthesis of polyunsaturated fatty acids (PUFAs, Figure 4), which might be one of the reasons for M. alpina CGMCC 20262 yielding more PUFAs than M. schmuckeri CGMCC 20261.
PUFA biosynthesis requires an activity of elongation of very long chain fatty acids proteins, as well as fatty acid desaturases such as delta-5, delta-6, delta-12, and omega-3 desaturase [12,16]. For example, the M. alpina 1S-4 mutant in delta-5 desaturase resulted in a huge accumulation of dihomo-γ-linolenic acid (DGLA), up to 43.3% of total fatty acids [70]. In the present study, the delta-5 desaturase of M. alpina is significantly different from other species in the family Mortierellaceae based on the phylogenetic analyses (Figure 6 and Supplementary File S2), which is probably one of the reasons for synthesizing high content of ARA [12,71,72]. Elongation of very long chain fatty acids proteins (ELOVLs) are involved in the long-chain polyunsaturated fatty acid synthetic pathway, and four gene types, ELOVL2, ELOVL3, ELOVL4, and ELOVL6, are identified herein in 19 strains of Mortierellaceae (Supplementary Figure S1 and Supplementary File S3).
The fatty acid synthase plays an important role in lipid synthesis, catalyzing the synthesis of saturated fatty acids from acetyl-CoA and malonyl-CoA. A variety of fatty acid synthases were found in fungi, such as Aspergillus oryzae, Mortierella alpina, Saccharomyces cerevisiae, and Yarrowia lipolytica [6,12,73,74]. In this study, three types of fatty acid synthases are found in Mortierellaceae (Supplementary Figure S2 and Supplementary File S1). Type I is the most common and similar to that reported in Mortierella alpina ATCC 32222, a single subunit consisting of eight catalytic domains/active sites [17]. Types II and III are new to science. Type II presents in M. antarctica KOD 1030 only, and type III in Dissophora globulifera REB-010B and Mortierella alpina AD071 (Supplementary Figure S2 and Supplementary File S1).
Overall, the rapid increase in genomic data is providing more materials for tackling fatty acid biosynthesis, discovering more potential genes, and finalizing the understanding of the genetic foundation of lipid metabolism.

5. Conclusions

In this paper, two new genomes, Mortierella alpina CGMCC 20262 and M. schmuckeri CGMCC 20261, were sequenced using the PacBio Sequel and Illumina NovaSeq 6000 platform. To explore the lipid metabolism in Mortierellaceae, a total of 19 genomes were reannotated. The results suggest that delta-5 desaturase and elongation of very long chain fatty acids protein 3 probably promoted the accumulation of polyunsaturated fatty acids, especially arachidonic acid. Besides, three types of fatty acid synthase, including two novel ones, were identified. Consequently, with the increase in public genomic data, a comprehensive analysis on lipid metabolism will discover more genes or protein-encoding models, providing more information for subsequent genetic engineering of oleaginous fungi.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/jof8090891/s1. Supplementary Table S1. The gene number of lipid metabolism in the 19 strains of Mortierellaceae; Supplementary Figure S1. A Maximum Likelihood tree of elongation of very long chain fatty acids protein in the family Mortierellaceae; Supplementary Figure S2. The heatmap of gene numbers associated with fatty acid synthesis annotated in 19 strains of Mortierellaceae; Supplementary File S1. The amino acid sequences of fatty acid synthase in Mortierellaceae; Supplementary File S2. The aligned dataset of amino acid sequences of delta-5 desaturase; Supplementary File S3. The amino acid sequences of elongation of very long chain fatty acids protein in Mortierellaceae.

Author Contributions

H.Z.; validation, writing—original draft preparation, and writing—review, Y.N.; formal analysis, Y.J.; and S.W.; investigation, T.-Y.Z.; data curation, X.-Y.L.; funding acquisition, projection administration, and writing—review. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Natural Science Foundation of China, Grant Nos. 31970009 and 32170012.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The newly identified genomics sequences have been deposited in GenBank (Table 2).

Acknowledgments

We thank Beijing Novogene Bioinformatics Technology Co., Ltd. (Beijing, China).

Conflicts of Interest

All authors declare no conflict of interest.

References

  1. Zhao, H.; Lv, M.L.; Liu, Z.; Zhang, M.Z.; Wang, Y.N.; Ju, X.; Song, Z.; Ren, L.Y.; Jia, B.S.; Qiao, M.; et al. High-yield oleaginous fungi and high-value microbial lipid resources from Mucoromycota. BioEnergy Res. 2021, 14, 1196–1206. [Google Scholar] [CrossRef]
  2. Reis, C.E.R.; Valle, G.F.; Bento, H.B.; Carvalho, A.K.; Alves, T.M.; de Castro, H.F. Sugarcane by-products within the biodiesel production chain: Vinasse and molasses as feedstock for oleaginous fungi and conversion to ethyl esters. Fuel 2020, 277, 118064. [Google Scholar] [CrossRef]
  3. Papanikolaou, S.; Aggelis, G. Sources of microbial oils with emphasis to Mortierella (Umbelopsis) isabellina fungus. World J. Microb. Biot. 2019, 35, 63. [Google Scholar] [CrossRef] [PubMed]
  4. Kothri, M.; Mavrommati, M.; Elazzazy, A.M.; Baeshen, M.N.; Moussa, T.A.; Aggelis, G. Microbial sources of polyunsaturated fatty acids (PUFAs) and the prospect of organic residues and wastes as growth media for PUFA-producing microorganisms. FEMS Microbiol. Lett. 2020, 367, fnaa028. [Google Scholar] [CrossRef] [PubMed]
  5. Blazeck, J.; Hill, A.; Liu, L.; Knight, R.; Miller, J.; Pan, A.; Otoupal, P.; Alper, H.S. Harnessing Yarrowia lipolytica lipogenesis to create a platform for lipid and biofuel production. Nat. Commun. 2014, 5, 3131. [Google Scholar] [CrossRef] [Green Version]
  6. Thammarongtham, C.; Nookaew, I.; Vorapreeda, T.; Srisuk, T.; Land, M.L.; Jeennor, S.; Laoteng, K. Genome characterization of oleaginous Aspergillus oryzae BCC7051: A potential fungal-based platform for lipid production. Curr. Microbiol. 2018, 75, 57–70. [Google Scholar] [CrossRef]
  7. Hashem, A.H.; Suleiman, W.B.; Abu-Elrish, G.M.; Ei-Sheikh, H.H. Consolidated bioprocessing of sugarcane bagasse to microbial oil by newly isolated oleaginous fungus: Mortierella wolfii. Arab. J. Sci. Eng. 2021, 46, 199–211. [Google Scholar] [CrossRef]
  8. Hashem, A.H.; Abu-Elreesh, G.; El-Sheikh, H.H.; Suleiman, W.B. Isolation, identification, and statistical optimization of a psychrotolerant Mucor racemosus for sustainable lipid production. Biomass Conv. Biorefinery 2022, 1–12. [Google Scholar] [CrossRef]
  9. Hashem, A.H.; Hasanin, M.S.; Khalil, A.M.A.; Suleiman, W.B. Eco-green conversion of watermelon peels to single cell oils using a unique oleaginous fungus: Lichtheimia corymbifera AH13. Waste Biomass Valor 2022, 11, 5721–5732. [Google Scholar] [CrossRef]
  10. Hashem, A.H.; Suleiman, W.B.; Abu-elreesh, G.; Shehabeldine, A.M.; Khalil, A.M.A. Sustainable lipid production from oleaginous fungus Syncephalastrum racemosum using synthetic and watermelon peel waste media. Bioresour. Technol. Rep. 2020, 12, 100569. [Google Scholar] [CrossRef]
  11. Gad, A.M.; Suleiman, W.B.; El-Sheikh, H.H.; Eimezayen, H.A.; Beltagy, A.M. Characterization of cellulase from Geotrichum candidum strain Gad1 approaching bioethanol production. Arab. J. Sci. Eng. 2022, 47, 6837–6850. [Google Scholar] [CrossRef]
  12. Wang, L.; Chen, W.; Feng, Y.; Ren, Y.; Gu, Z.; Chen, H.; Wang, H.; Thomas, M.J.; Zhang, B.; Berquin, I.M. Genome characterization of the oleaginous fungus Mortierella alpina. PLoS ONE 2011, 6, e28319. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Bánki, O.; Roskov, Y.; Döring, M.; Ower, G.; Vandepitte, L.; Hobern, D.; Remsen, D.; Schalk, P.; DeWalt, R.; Keping, M.; et al. Catalogue of Life Checklist, Version 2022-03-21; Catalogue of Life. 2022. Available online: https://www.catalogueoflife.org/ (accessed on 2 July 2022).
  14. Eroshin, V.; Dedyukhina, E.; Chistyakova, T.; Zhelifonova, V.; Kurtzman, C.; Bothast, R. Arachidonic-acid production by species of Mortierella. World J. Microbiol. Biotechnol. 1996, 12, 91–96. [Google Scholar] [CrossRef] [PubMed]
  15. Botha, A.; Paul, I.; Roux, C.; Kock, J.L.; Coetzee, D.J.; Strauss, T.; Maree, C. An isolation procedure for arachidonic acid producing Mortierella species. Antonie. Leeuw. 1999, 75, 253–256. [Google Scholar] [CrossRef]
  16. Kikukawa, H.; Sakuradani, E.; Ando, A.; Shimizu, S.; Ogawa, J. Arachidonic acid production by the oleaginous fungus Mortierella alpina 1S-4: A review. J. Adv. Res. 2018, 11, 15–22. [Google Scholar] [CrossRef]
  17. Ishikura, Y.; Ikeda, G.; Akimoto, K.; Hata, M.; Kusumoto, A.; Kidokoro, A.; Kontani, M.; Kawashima, H.; Kiso, Y.; Koga, Y. Arachidonic acid supplementation decreases P300 latency and increases P300 amplitude of event-related potentials in healthy elderly men. Neuropsychobiology 2009, 60, 73–79. [Google Scholar] [CrossRef]
  18. Tallima, H.; El Ridi, R. Arachidonic acid: Physiological roles and potential health benefits–a review. J. Adv. Res. 2018, 11, 33–41. [Google Scholar] [CrossRef]
  19. Ratledge, C.; Wynn, J.P. The biochemistry and molecular biology of lipid accumulation in oleaginous microorganisms. In Advances in Applied Microbiology; Academic Press: San Diego, CA, USA, 2002; Volume 51, pp. 1–51. [Google Scholar]
  20. Ho, S.-Y.; Jiang, Y.; Chen, F. Polyunsaturated fatty acids (PUFAs) content of the fungus Mortierella alpina isolated from soil. J. Agric. Food Chem. 2007, 55, 3960–3966. [Google Scholar] [CrossRef]
  21. Tedersoo, L.; Sánchez-Ramírez, S.; Kõljalg, U.; Bahram, M.; Döring, M.; Schigel, D.; May, T.; Ryberg, M.; Abarenkov, K. High-level classification of the Fungi and a tool for evolutionary ecological analyses. Fungal Divers. 2018, 90, 135–159. [Google Scholar] [CrossRef] [Green Version]
  22. Kikukawa, H.; Sakuradani, E.; Ando, A.; Okuda, T.; Shimizu, S.; Ogawa, J. Microbial production of dihomo-γ-linolenic acid by Δ5-desaturase gene-disruptants of Mortierella alpina 1S-4. J. Biosci. Bioeng. 2016, 122, 22–26. [Google Scholar] [CrossRef] [Green Version]
  23. Chang, L.; Tang, X.; Lu, H.; Zhang, H.; Chen, Y.Q.; Chen, H.; Chen, W. Role of adenosine monophosphate deaminase during fatty acid accumulation in oleaginous fungus Mortierella alpina. J. Agric. Food Chem. 2019, 67, 9551–9559. [Google Scholar] [CrossRef] [PubMed]
  24. Rong, C.; Chen, H.; Tang, X.; Gu, Z.; Zhao, J.; Zhang, H.; Chen, Y.; Chen, W. Structural determinants of substrate specificity of omega-3 desaturases from Mortierella alpina and Rhizophagus irregularis by domain-swapping and molecular docking. Int. J. Mol. Sci. 2019, 20, 1603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Vorapreeda, T.; Thammarongtham, C.; Palasak, T.; Srisuk, T.; Jenjaroenpun, P.; Wongsurawat, T.; Nookaew, I.; Laoteng, K. Systematic genome analysis of a novel arachidonic acid-producing strain uncovered unique metabolic traits in the production of acetyl-CoA-derived products in Mortierella fungi. Gene 2020, 741, 144559. [Google Scholar] [CrossRef] [PubMed]
  26. Etienne, K.A.; Chibucos, M.C.; Su, Q.; Orvis, J.; Daugherty, S.; Ott, S.; Sengamalay, N.A.; Fraser, C.M.; Lockhart, S.R.; Bruno, V.M. Draft genome sequence of Mortierella alpina isolate CDC-B6842. Genome Announc. 2014, 2, e01180-13. [Google Scholar] [CrossRef] [Green Version]
  27. Vandepol, N.; Liber, J.; Desirò, A.; Na, H.; Kennedy, M.; Barry, K.; Grigoriev, I.V.; Miller, A.N.; O’Donnell, K.; Stajich, J.E. Resolving the Mortierellaceae phylogeny through synthesis of multi-gene phylogenetics and phylogenomics. Fungal Divers. 2020, 104, 267–289. [Google Scholar] [CrossRef]
  28. Yang, S.; Vinatzer, B.A. Draft Genome sequence of Mortierella alpina Strain LL118, isolated from an Aspen (Populus tremuloides) leaf litter sample. Microbiol. Resour. Announc. 2021, 10, e00864-21. [Google Scholar] [CrossRef]
  29. Seif, E.; Leigh, J.; Liu, Y.; Roewer, I.; Forget, L.; Lang, B.F. Comparative mitochondrial genomics in zygomycetes: Bacteria-like RNase P RNAs, mobile elements and a close source of the group I intron invasion in angiosperms. Nucleic Acids Res. 2005, 33, 734–744. [Google Scholar] [CrossRef] [Green Version]
  30. Uehling, J.; Gryganskyi, A.; Hameed, K.; Tschaplinski, T.; Misztal, P.; Wu, S.; Desirò, A.; Vande Pol, N.; Du, Z.; Zienkiewicz, A. Comparative genomics of Mortierella elongata and its bacterial endosymbiont Mycoavidus cysteinexigens. Environ. Microbiol. 2017, 19, 2964–2983. [Google Scholar] [CrossRef]
  31. Yang, Y.; Liu, X.Y.; Huang, B. The complete mitochondrial genome of Linnemannia amoeboidea (W. Gams) Vandepol & Bonito (Mortierellales: Mortierellaceae). Mitochondrial DNA B 2022, 7, 374–376. [Google Scholar] [CrossRef]
  32. Kendrick, A.; Ratledge, C. Desaturation of polyunsaturated fatty acids in Mucor circinelloides and the involvement of a novel membrane-bound malic enzyme. Eur. J. Biochem. 1992, 209, 667–673. [Google Scholar] [CrossRef]
  33. Betina, V.; Koman, V. Changes in the lipid composition during the photo-induced conidiation of Trichoderma viride. Folia Microbiol. 1980, 25, 295. [Google Scholar] [CrossRef] [PubMed]
  34. Chin, C.S.; Alexander, D.H.; Marks, P.; Klammer, A.A.; Drake, J.; Heiner, C.; Clum, A.; Copeland, A.; Huddleston, J.; Eichler, E.E. Nonhybrid, finished microbial genome assemblies from long-read SMRT sequencing data. Nat. Methods 2013, 10, 563–569. [Google Scholar] [CrossRef] [PubMed]
  35. Gurevich, A.; Saveliev, V.; Vyahhi, N.; Tesler, G. QUAST: Quality assessment tool for genome assemblies. Bioinformatics 2013, 29, 1072–1075. [Google Scholar] [CrossRef]
  36. Simão, F.A.; Waterhouse, R.M.; Ioannidis, P.; Kriventseva, E.V.; Zdobnov, E.M. BUSCO: Assessing genome assembly and annotation completeness with single-copy orthologs. Bioinformatics 2015, 31, 3210–3212. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Stanke, M.; Waack, S. Gene prediction with a hidden Markov model and a new intron submodel. Bioinformatics 2003, 19, ii215–ii225. [Google Scholar] [CrossRef] [Green Version]
  38. El-Gebali, S.; Mistry, J.; Bateman, A.; Eddy, S.R.; Luciani, A.; Potter, S.C.; Qureshi, M.; Richardson, L.J.; Salazar, G.A.; Smart, A. The Pfam protein families database in 2019. Nucleic Acids Res. 2019, 47, D427–D432. [Google Scholar] [CrossRef]
  39. Ashburner, M.; Ball, C.A.; Blake, J.A.; Botstein, D.; Butler, H.; Cherry, J.M.; Davis, A.P.; Dolinski, K.; Dwight, S.S.; Eppig, J.T. Gene ontology: Tool for the unification of biology. Nat. Genet. 2000, 25, 25–29. [Google Scholar] [CrossRef] [Green Version]
  40. Kanehisa, M.; Goto, S.; Hattori, M.; Aoki-Kinoshita, K.F.; Itoh, M.; Kawashima, S.; Katayama, T.; Araki, M.; Hirakawa, M. From genomics to chemical genomics: New developments in KEGG. Nucleic Acids Res. 2006, 34, D354–D357. [Google Scholar] [CrossRef]
  41. Cantarel, B.L.; Coutinho, P.M.; Rancurel, C.; Bernard, T.; Lombard, V.; Henrissat, B. The Carbohydrate-Active EnZymes database (CAZy): An expert resource for glycogenomics. Nucleic Acids Res. 2009, 37, D233–D238. [Google Scholar] [CrossRef]
  42. Zhang, H.; Yohe, T.; Entwustke, S.; Wu, P.Z.; Yang, Z.L.; Busk, P.K.; Xu, Y.; Yin, Y.B. dbCAN2: A meta server for automated carbohydrate-activate enzyme annotation. Nucleic Acids Res. 2018, 46, W95–W101. [Google Scholar] [CrossRef] [Green Version]
  43. Huerta-Cepas, J.; Szklarczyk, D.; Heller, D.; Hernández-Plaza, A.; Forslund, S.K.; Cook, H.; Mende, D.R.; Letunic, I.; Rattei, T.; Jensen, L.J. eggNOG 5.0: A hierarchical, functionally and phylogenetically annotated orthology resource based on 5090 organisms and 2502 viruses. Nucleic Acids Res. 2019, 47, D309–D314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Buchfink, B.; Xie, C.; Huson, D.H. Fast and sensitive protein alignment using DIAMOND. Nat. Methods 2015, 12, 59–60. [Google Scholar] [CrossRef] [PubMed]
  45. Medema, M.H.; Blin, K.; Cimermancic, P.; De Jager, V.; Zakrzewski, P.; Fischbach, M.A.; Weber, T.; Takano, E.; Breitling, R. antiSMASH: Rapid identification, annotation and analysis of secondary metabolite biosynthesis gene clusters in bacterial and fungal genome sequences. Nucleic Acids Res. 2011, 39, W339–W346. [Google Scholar] [CrossRef] [PubMed]
  46. Ou, S.; Su, W.; Liao, Y.; Chougule, K.; Agda, J.R.; Hellinga, A.J.; Lugo, C.S.B.; Elliott, T.A.; Ware, D.; Peterson, T. Benchmarking transposable element annotation methods for creation of a streamlined, comprehensive pipeline. Genome Biol. 2019, 20, 275. [Google Scholar] [CrossRef] [Green Version]
  47. Lagesen, K.; Hallin, P.; Rødland, E.A.; Stærfeldt, H.H.; Rognes, T.; Ussery, D.W. RNAmmer: Consistent and rapid annotation of ribosomal RNA genes. Nucleic Acids Res. 2007, 35, 3100–3108. [Google Scholar] [CrossRef]
  48. Lowe, T.M.; Eddy, S.R. tRNAscan-SE: A program for improved detection of transfer RNA genes in genomic sequence. Nucleic Acids Res. 1997, 25, 955–964. [Google Scholar] [CrossRef]
  49. Finn, R.D.; Clements, J.; Eddy, S.R. HMMER web server: Interactive sequence similarity searching. Nucleic Acids Res. 2011, 39, W29–W37. [Google Scholar] [CrossRef] [Green Version]
  50. Sánchez, R.; Serra, F.; Tárraga, J.; Medina, I.; Carbonell, J.; Pulido, L.; de María, A.; Capella-Gutíerrez, S.; Huerta-Cepas, J.; Gabaldón, T. Phylemon 2.0: A suite of web-tools for molecular evolution, phylogenetics, phylogenomics and hypotheses testing. Nucleic Acids Res. 2011, 39, W470–W474. [Google Scholar] [CrossRef] [Green Version]
  51. Spatafora, J.W.; Chang, Y.; Benny, G.L.; Lazarus, K.; Smith, M.E.; Berbee, M.L.; Bonito, G.; Corradi, N.; Grigoriev, I.; Gryganskyi, A. A phylum-level phylogenetic classification of zygomycete fungi based on genome-scale data. Mycologia 2016, 108, 1028–1046. [Google Scholar] [CrossRef] [Green Version]
  52. James, T.Y.; Pelin, A.; Bonen, L.; Ahrendt, S.; Sain, D.; Corradi, N.; Stajich, J.E. Shared signatures of parasitism and phylogenomics unite Cryptomycota and microsporidia. Curr. Biol. 2013, 23, 1548–1553. [Google Scholar] [CrossRef] [Green Version]
  53. Katoh, K.; Standley, D.M. MAFFT multiple sequence alignment software version 7: Improvements in performance and usability. Mol. Biol. Evol. 2013, 30, 772–780. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Stamatakis, A. RAxML version 8: A tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 2014, 30, 1312–1313. [Google Scholar] [CrossRef] [PubMed]
  55. Nguyen, L.-T.; Schmidt, H.A.; Von Haeseler, A.; Minh, B.Q. IQ-TREE: A fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Mol. Biol. Evol. 2015, 32, 268–274. [Google Scholar] [CrossRef] [PubMed]
  56. Ellenberger, S.; Burmester, A.; Wöstemeyer, J. Complete mitochondrial DNA sequence of the mucoralean fungus Absidia glauca, a model for studying host-parasite interactions. Genome Announc. 2016, 4, e00153-16. [Google Scholar] [CrossRef] [Green Version]
  57. Chang, Y.; Desirò, A.; Na, H.; Sandor, L.; Lipzen, A.; Clum, A.; Barry, K.; Grigoriev, I.V.; Martin, F.M.; Stajich, J.E. Phylogenomics of Endogonaceae and evolution of mycorrhizas within Mucoromycota. New Phytol. 2019, 222, 511–525. [Google Scholar] [CrossRef] [Green Version]
  58. Schwartze, V.U.; Winter, S.; Shelest, E.; Marcet-Houben, M.; Horn, F.; Wehner, S.; Linde, J.; Valiante, V.; Sammeth, M.; Riege, K. Gene expansion shapes genome architecture in the human pathogen Lichtheimia corymbifera: An evolutionary genomics analysis in the ancient terrestrial Mucorales (Mucoromycotina). PLoS Genet. 2014, 10, e1004496. [Google Scholar] [CrossRef]
  59. Linde, J.; Schwartze, V.; Binder, U.; Lass-Flörl, C.; Voigt, K.; Horn, F. De novo whole-genome sequence and genome annotation of Lichtheimia ramosa. Genome Announc. 2014, 2, e00888-14. [Google Scholar] [CrossRef] [Green Version]
  60. Lee, S.C.; Billmyre, R.B.; Li, A.; Carson, S.; Sykes, S.M.; Huh, E.Y.; Mieczkowski, P.; Ko, D.C.; Cuomo, C.A.; Heitman, J. Analysis of a food-borne fungal pathogen outbreak: Virulence and genome of a Mucor circinelloides isolate from yogurt. MBio 2014, 5, e01390-14. [Google Scholar] [CrossRef] [Green Version]
  61. Corrochano, L.M.; Kuo, A.; Marcet-Houben, M.; Polaino, S.; Salamov, A.; Villalobos-Escobedo, J.M.; Grimwood, J.; Álvarez, M.I.; Avalos, J.; Bauer, D. Expansion of signal transduction pathways in fungi by extensive genome duplication. Curr. Biol. 2016, 26, 1577–1584. [Google Scholar] [CrossRef] [Green Version]
  62. Nguyen, M.H.; Kaul, D.; Muto, C.; Cheng, S.J.; Richter, R.A.; Bruno, V.M.; Liu, G.; Beyhan, S.; Sundermann, A.J.; Mounaud, S. Genetic diversity of clinical and environmental Mucorales isolates obtained from an investigation of mucormycosis cases among solid organ transplant recipients. Microb. Genom. 2020, 6, mgen000473. [Google Scholar] [CrossRef]
  63. Horn, F.; Üzüm, Z.; Möbius, N.; Guthke, R.; Linde, J.; Hertweck, C. Draft genome sequences of symbiotic and nonsymbiotic Rhizopus microsporus strains CBS 344.29 and ATCC 62417. Genome Announc. 2015, 3, e01370-14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Wang, Y.; White, M.M.; Kvist, S.; Moncalvo, J.-M. Genome-wide survey of gut fungi (Harpellales) reveals the first horizontally transferred ubiquitin gene from a mosquito host. Mol. Biol. Evol. 2016, 33, 2544–2554. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Takeda, I.; Tamano, K.; Yamane, N.; Ishii, T.; Miura, A.; Umemura, M.; Terai, G.; Baker, S.E.; Koike, H.; Machida, M. Genome sequence of the Mucoromycotina fungus Umbelopsis isabellina, an effective producer of lipids. Genome Announc. 2014, 2, e00071-14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Wang, Y.; Tang, H.; DeBarry, J.D.; Tan, X.; Li, J.; Wang, X.; Lee, T.H.; Jin, H.; Marler, B.; Guo, H. MCScanX: A toolkit for detection and evolutionary analysis of gene synteny and collinearity. Nucleic Acids Res. 2012, 40, e49. [Google Scholar] [CrossRef] [Green Version]
  67. Dulermo, T.; Nicaud, J.-M. Involvement of the G3P shuttle and β-oxidation pathway in the control of TAG synthesis and lipid accumulation in Yarrowia lipolytica. Metab. Eng. 2011, 13, 482–491. [Google Scholar] [CrossRef]
  68. Beopoulos, A.; Cescut, J.; Haddouche, R.; Uribelarrea, J.-L.; Molina-Jouve, C.; Nicaud, J.M. Yarrowia lipolytica as a model for bio-oil production. Prog. Lipid Res. 2009, 48, 375–387. [Google Scholar] [CrossRef]
  69. Rodríguez-Frómeta, R.A.; Gutiérrez, A.; Torres-Martínez, S.; Garre, V. Malic enzyme activity is not the only bottleneck for lipid accumulation in the oleaginous fungus Mucor circinelloides. Appl. Microbiol. Biot. 2013, 97, 3063–3072. [Google Scholar] [CrossRef]
  70. Jareonkitmongkol, S.; Sakuradani, E.; Shimizu, S. A novel Δ5-desaturase-defective mutant of Mortierella alpina 1S-4 and its dihomo-γ-linolenic acid productivity. Appl. Environ. Microb. 1993, 59, 4300–4304. [Google Scholar] [CrossRef] [Green Version]
  71. Parker-Barnes, J.M.; Das, T.; Bobik, E.; Leonard, A.E.; Thurmond, J.M.; Chaung, L.T.; Huang, Y.S.; Mukerji, P. Identification and characterization of an enzyme involved in the elongation of n-6 and n-3 polyunsaturated fatty acids. Proc. Natl. Acad. Sci. USA 2000, 97, 8284–8289. [Google Scholar] [CrossRef] [Green Version]
  72. Sakuradani, E.; Nojiri, M.; Suzuki, H.; Shimizu, S. Identification of a novel fatty acid elongase with a wide substrate specificity from arachidonic acid-producing fungus Mortierella alpina 1S-4. Appl. Microbiol. Biot. 2009, 84, 709–716. [Google Scholar] [CrossRef]
  73. Shin, G.H.; Veen, M.; Stahl, U.; Lang, C. Overexpression of genes of the fatty acid biosynthetic pathway leads to accumulation of sterols in Saccharomyces cerevisiae. Yeast 2012, 29, 371–383. [Google Scholar] [CrossRef] [PubMed]
  74. Tamano, K.; Bruno, K.S.; Karagiosis, S.A.; Culley, D.E.; Deng, S.; Collett, J.R.; Umemura, M.; Koike, H.; Baker, S.E.; Machida, M. Increased production of fatty acids and triglycerides in Aspergillus oryzae by enhancing expressions of fatty acid synthesis-related genes. Appl. Microbiol. Biot. 2013, 97, 269–281. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Lipids of Mortierella alpina CGMCC 20262 and M. schmuckeri CGMCC 20261. (a) Lipid characters: dry cell weight (DCW, g/L), total lipid weight (TLW, g/L), and total lipid content (TLC); (b) fatty acid profiles: others include C10:0, C12:0, C15:0, C20:0, and C20:1.
Figure 1. Lipids of Mortierella alpina CGMCC 20262 and M. schmuckeri CGMCC 20261. (a) Lipid characters: dry cell weight (DCW, g/L), total lipid weight (TLW, g/L), and total lipid content (TLC); (b) fatty acid profiles: others include C10:0, C12:0, C15:0, C20:0, and C20:1.
Jof 08 00891 g001
Figure 2. A Maximum Likelihood phylogenomic tree illustrating the placements of Mortierella alpina CGMCC 20262 and M. schmuckeri CGMCC 20261 based on 192 clusters of orthologous proteins. New genomes obtained in this study are in bold. Maximum Likelihood bootstrap values (MLBV ≥ 50%) are indicated along branches. A scale bar in the upper left indicates substitutions per site.
Figure 2. A Maximum Likelihood phylogenomic tree illustrating the placements of Mortierella alpina CGMCC 20262 and M. schmuckeri CGMCC 20261 based on 192 clusters of orthologous proteins. New genomes obtained in this study are in bold. Maximum Likelihood bootstrap values (MLBV ≥ 50%) are indicated along branches. A scale bar in the upper left indicates substitutions per site.
Jof 08 00891 g002
Figure 3. Genomic synteny of Mortierella alpina CGMCC 20262 and M. schmuckeri CGMCC 20261 based on protein-coding gene models of the 10 largest scaffolds.
Figure 3. Genomic synteny of Mortierella alpina CGMCC 20262 and M. schmuckeri CGMCC 20261 based on protein-coding gene models of the 10 largest scaffolds.
Jof 08 00891 g003
Figure 4. The polyunsaturated fatty acid synthesis mechanism in Mortierellaceae. EMP: Glycolytic pathway; TCA cycle: Tricarboxylic acid cycle; ME: Malate dehydrogenase; NADPH: Nicotinamide adenine dinucleotide phosphate; ACC: Acetyl-CoA carboxylase; ACL: ATP citrate (pro-S)-lyase; FAS: Fatty acid synthase; ELO2: Fatty acid elongase 2; Delta5: delta-5 desaturase; Delta6: delta-6 desaturase; Delta9: delta-9 desaturase; Delta12: delta-12 desaturase; ELOVL: Elongation of very long chain fatty acids protein.
Figure 4. The polyunsaturated fatty acid synthesis mechanism in Mortierellaceae. EMP: Glycolytic pathway; TCA cycle: Tricarboxylic acid cycle; ME: Malate dehydrogenase; NADPH: Nicotinamide adenine dinucleotide phosphate; ACC: Acetyl-CoA carboxylase; ACL: ATP citrate (pro-S)-lyase; FAS: Fatty acid synthase; ELO2: Fatty acid elongase 2; Delta5: delta-5 desaturase; Delta6: delta-6 desaturase; Delta9: delta-9 desaturase; Delta12: delta-12 desaturase; ELOVL: Elongation of very long chain fatty acids protein.
Jof 08 00891 g004
Figure 5. Fatty acid synthases in Mortierellaceae. AT: Acyl transferase domain; ER: Enoyl reductase domain; DH: dehydratase; MPT: Malonyl CoA-acyl carrier protein transacylase; ACP: Acyl carrier protein domain; KR: β-ketoacyl reductase; KS: β-ketoacyl synthase; PPT: Phosphopantetheinyl transferase.
Figure 5. Fatty acid synthases in Mortierellaceae. AT: Acyl transferase domain; ER: Enoyl reductase domain; DH: dehydratase; MPT: Malonyl CoA-acyl carrier protein transacylase; ACP: Acyl carrier protein domain; KR: β-ketoacyl reductase; KS: β-ketoacyl synthase; PPT: Phosphopantetheinyl transferase.
Jof 08 00891 g005
Figure 6. A Maximum Likelihood phylogenetic tree of the gene encoding delta-5 desaturase in the family Mortierellaceae. A scale bar in the upper left indicates substitutions per site.
Figure 6. A Maximum Likelihood phylogenetic tree of the gene encoding delta-5 desaturase in the family Mortierellaceae. A scale bar in the upper left indicates substitutions per site.
Jof 08 00891 g006
Table 1. The genomic information used for phylogenomic analyses in this study.
Table 1. The genomic information used for phylogenomic analyses in this study.
SpeciesStrainsBioSampleReferences
Absidia glaucaCBS 101.48SAMEA3923633[56]
Acaulospora morrowiaeCL551SAMEA8911292
Apophysomyces elegansB7760SAMN02351510
Actinomortierella wolfii *NRRL 6351SAMN05720777[27]
Choanephora cucurbitarumKUS-F282377SAMN04532838
Cunninghamella elegansB9769SAMN02351511
Dissophora globulifera *REB 010BSAMN05720531[27]
Diversispora eburneaAZ414ASAMEA8911293
Endogone sp.FLAS-F59071SAMN09071421[57]
Gongronella butleriGbKAUSAMN15221701
Lichtheimia corymbiferaFSU 9682SAMEA2189700[58]
L. ramosaFSU 6197SAMN05179542[59]
Mortierella alpina *CGMCC 20262SAMN29490473This study
M. alpina *AD071SAMN05720461[27]
M. alpina *AD072SAMN05720462[27]
M. alpina *ATCC 32222SAMN02981246[12]
M. alpina *B6842SAMN02370960[26]
M. alpina *CCTCC M-207067SAMN03658567
M. alpina *CK1249SAMN05720518[27]
M. alpina *GBA31SAMN05720773[27]
M. alpina *LL118SAMN20056918[28]
M. alpina *NRRL 66262SAMN10361219[27]
M. amoeboidea *CBS 889.72SAMN19911466[31]
M. antarctica *KOD1030SAMN05720520[27]
M. elongata *AG-77SAMN02745706[30]
M. epicladia *AD058SAMN05720441[27]
M. gamsii *NVP60SAMN05720530[27]
M. verticillata *NRRL 6337SAMN00699802[29]
M. schmuckeri *CGMCC 20261SAMN29492047This study
Mucor circinelloides1006PhLSAMN00103456[60]
M. lusitanicusCBS 277.49SAMN00120579[61]
Parasitella parasiticaCBS 412.66SAMEA278055[56]
Rhizopus arrhizusGL8SAMN14162349[62]
R. microsporusATCC 52813SAMN06821222[63]
Phycomyces blakesleeanusNRRL 1555SAMN00189023[61]
Smittium culicisGSMNPSAMN04489870[64]
Syncephalastrum monosporumB8922SAMN02370995
Thermomucor indicae-seudaticaeHACC 243SAMN03070115
Umbelopsis isabellinaWA0000067209SAMN16393839[65]
U. vinaceaWA0000051536SAMN16393840
Note: The star mark “*” represents members in the Mortierellaceae. The two genomes newly generated in this study are in bold.
Table 2. Genomic features of Mortierella sequenced and de novo assembled in this study.
Table 2. Genomic features of Mortierella sequenced and de novo assembled in this study.
Species M. alpina CGMCC 20262M. schmuckeri CGMCC 20261
Genome size (Mb) 40.2249.24
Scaffolds 2825
Largest scaffolds (Mb) 4.354.78
GC (%) 50.9247.46
N50 (Mb) 2.532.71
L50 68
Assembly BUSCO coverage (%) 97.497.6
PCG models 11,76113,051
Pfam85078636
NT40264434
NR49046180
GO54115571
Eggnog90769436
KEGG48584921
CAZymes229222
Gene clusters of secondary metabolites
Terpene42
Fungal-RiPP01
NRPS150
NRPS-like12
Siderophore20
Repetitive elements (% in genomes) 7.297.45
ncRNA
rRNA3829
tRNA226262
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhao, H.; Nie, Y.; Jiang, Y.; Wang, S.; Zhang, T.-Y.; Liu, X.-Y. Comparative Genomics of Mortierellaceae Provides Insights into Lipid Metabolism: Two Novel Types of Fatty Acid Synthase. J. Fungi 2022, 8, 891. https://doi.org/10.3390/jof8090891

AMA Style

Zhao H, Nie Y, Jiang Y, Wang S, Zhang T-Y, Liu X-Y. Comparative Genomics of Mortierellaceae Provides Insights into Lipid Metabolism: Two Novel Types of Fatty Acid Synthase. Journal of Fungi. 2022; 8(9):891. https://doi.org/10.3390/jof8090891

Chicago/Turabian Style

Zhao, Heng, Yong Nie, Yang Jiang, Shi Wang, Tian-Yu Zhang, and Xiao-Yong Liu. 2022. "Comparative Genomics of Mortierellaceae Provides Insights into Lipid Metabolism: Two Novel Types of Fatty Acid Synthase" Journal of Fungi 8, no. 9: 891. https://doi.org/10.3390/jof8090891

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop