Next Article in Journal
Epigenetic Mechanisms Impacting Aging: A Focus on Histone Levels and Telomeres
Next Article in Special Issue
Patchoulol Production with Metabolically Engineered Corynebacterium glutamicum
Previous Article in Journal
Alteration of Hepatic Gene Expression along with the Inherited Phenotype of Acquired Fatty Liver in Chicken
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Possible Trifunctional β-Carotene Synthase Gene Identified in the Draft Genome of Aurantiochytrium sp. Strain KH105

1
Department of Molecular Biotechnology, Graduate School of Advanced Sciences of Matter, Hiroshima University, Higashi-Hiroshima 739-8530, Japan
2
Marine Genomic Unit, Okinawa Institute of Science and Technology Graduate University, Onna, Okinawa 904-0945, Japan
3
JST-CREST, Saitama 332-0012, Japan
*
Authors to whom correspondence should be addressed.
These authors contributed equally to the work.
Submission received: 14 February 2018 / Revised: 27 March 2018 / Accepted: 27 March 2018 / Published: 9 April 2018
(This article belongs to the Special Issue Microbial Metabolic Engineering)

Abstract

:
Labyrinthulomycetes have been regarded as a promising industrial source of xanthophylls, including astaxanthin and canthaxanthin, polyunsaturated fatty acids such as docosahexaenoic acid and docosapentaenoic acid, ω-3 oils, and terpenic hydrocarbons, such as sterols and squalene. A Thraustochytrid, Aurantiochytrium sp. KH105 produces carotenoids, including astaxanthin, with strong antioxidant activity. To gain genomic insights into this capacity, we decoded its 97-Mbp genome and characterized genes for enzymes involved in carotenoid biosynthesis. Interestingly, all carotenogenic genes, as well as other eukaryotic genes, appeared duplicated, suggesting that this strain is diploid. In addition, among the five genes involved in the pathway from geranylgeranyl pyrophosphate to astaxanthin, geranylgeranyl phytoene synthase (crtB), phytoene desaturase (crtI) and lycopene cyclase (crtY) were fused into single gene (crtIBY) with no internal stop codons. Functionality of the trifunctional enzyme, CrtIBY, to catalyze the reaction from geranylgeranyl diphosphate to β-carotene was confirmed using a yeast assay system and mass spectrometry. Furthermore, analyses of differential gene expression showed characteristic up-regulation of carotenoid biosynthetic genes during stationary and starvation phases under these culture conditions. This suggests genetic engineering events to promote more efficient production of carotenoids. We also showed an occurrence of crtIBY in other Thraustochytrid species.

1. Introduction

A marine microbe, Aurantiochytrium, belongs to the order Thraustochytriaceae (so-called Thraustochytrids), pertaining to the class Labyrinthulomycetes, of the phylum Heterokonta, within the kingdom Chromista [1,2]. Thraustochytrids are a group of non-photosynthetic, marine, fungoid protists, characterized by an ectoplasmic net and a cell wall composed of non-cellulosic, sulfated scales [1,2,3]. It has been argued that the biomass of Labyrinthulomycetes in the water column may equal or even exceed that of bacteria [4]. Their high biomass and production of degradative enzymes indicate that Labyrinthulomycetes play a significant ecological role as alternative food sources for picoplankton feeders, as well as being active decomposers and consumers in marine microbial food chains [5].
Labyrinthulomycetes have also been regarded as a promising industrial source of xanthophylls, including astaxanthin and canthaxanthin, polyunsaturated fatty acids (PUFA) such as docosahexaenoic acid (DHA; 22:6n-3) and docosapentaenoic acid (DPA; 22:5n-6), ω-3 oils [6], and terpenic hydrocarbons, such as sterols and squalene [7,8,9]. In a previous study, we isolated and characterized Aurantiochytrium sp. strain KH105 as a potential source of carotenoids [10]. This strain produces β-carotene as well as xanthophylls such as echinenone, canthaxanthin, phoenicoxanthin, and astaxanthin [10,11]. The intracellular composition of these carotenoids is affected by media and culture conditions. However, the biosynthetic mechanism of these carotenoids has not been elaborated. Recently, genomes of several Thraustochytrid species have been decoded [12,13,14,15]. Here, we decoded the genome of Aurantiochytrium sp. strain KH105 specifically to understand its capacity to synthesize carotenoids.

2. Materials and Methods

2.1. Microorganism

Aurantiochytrium sp. strain KH105 [11], formerly classified in the genus Schizochytrium, but amended later [3], was maintained on GPY agar medium composed of 3% glucose, 0.6% hipolypeptone (Nihon Pharmaceutical, Tokyo, Japan), 0.2% yeast extract (Kyokuto Pharmaceutical Industrial, Tokyo, Japan), and 2% sea salts (Merck KGaA, Darmstadt, Germany), pH 7.0, with 1.8% agar at 28 °C, and pre-cultured in GPY liquid medium at 300 rpm for up to 24 h. A baffled 200-mL flask with 50 mL of GPY medium was inoculated with 1 mL of pre-cultured medium and cultured at 28 °C, 160 rpm. DNA extraction was performed on cells incubated for 96 h at 28 °C, and RNA extraction from cells cultured at 28 °C for 27 h (mid-log phase), 40 h (late-log to stationary phase) and 48 h (starvation phase), respectively.

2.2. DNA Preparation

Cells were harvested from culture broth by centrifugation (10,000× g, 20 min at 4 °C), and frozen in liquid nitrogen. Nuclear DNA for sequencing was obtained using a protease, RNAase, and a phenol-chloroform extraction protocol [16]. Packed, frozen cells were pulverized using a sterile mortar and pestle in liquid nitrogen. The powder was treated with 2% cetyltrimethylammonium bromide (CTAB) solution (2% CTAB, 100 mM Tris-HCl, pH 8.0, 20 mM ethylenediaminetetraacetic acid (EDTA), 1.4 M NaCl) for 1 h at 65 °C. Solutions were centrifuged at 1500× g for 5 min at 4 °C. Supernatant was subjected to 20 μg/mL RNase treatment (37 °C, 30 min) and further treated by adding an equal volume of chloroform/isoamyl alcohol (24:1, v/v) with gentle rotation for 1 h. After centrifugation at 1500× g using Phase lock gel tubes (Funakoshi, Tokyo, Japan), supernatants were transferred to new tubes. An additional chloroform treatment was repeated twice. By adding 1.5× volumes of 1% CTAB, genomic DNA was precipitated. Precipitated DNA was treated with 0.1 mg/mL proteinase K (Takara Bio Inc., Shiga, Japan) in 1 M NaCl for 2–3 h at 56 °C. By adding at least 2.5× volumes of ethanol, DNA was precipitated again. After rinsing with 70% ethanol twice, DNA was dissolved overnight in 10 mM Tris-HCl, 1 mM EDTA, pH 7.4, at 4 °C.

2.3. Genome Sequencing and Assembly

Genomic DNA was fragmented, and libraries were prepared and sequenced according to the manufacturer’s protocols with whole-genome shotgun (WGS) reads using a Roche 454 GS-FLX (Roche Diagnostics, Basel, Switzerland) [17], and Illumina Miseq and GAIIx sequencers (Illumina, San Diego, CA, USA) [18]. To obtain contigs, 454 and Miseq paired-end reads (Illumina, San Diego, CA, USA) were combined and assembled de novo using GS De Novo Assembler version 2.6 (Newbler, Roche, Basel, Switzerland) [17]. Then, subsequent scaffolding was performed with SOPRA [19] and SSPACE [20] using Illumina mate-pair information. At least five consistent read-pairs were needed to form a connection. Gaps inside scaffolds were closed with Illumina paired-end data using GapCloser [21]. Genome size can be roughly estimated by the k-mer value of new-generation sequence reads. Illumina sequences were quality trimmed (QV ≥ 20) and possible polymerase chain reaction (PCR) duplicates were removed with MarkDuplicates in Picard tools (http://picard.sourceforge.net, Broad Institute, Cambridge, MA, USA). K-mers (length: 21, 31, 41, 51, 61, and 71) in the dataset were counted using JELLYFISH [22].

2.4. RNA Sequencing Analyses

We performed transcriptome sequencing using an Illumina GAIIx sequencer (Illumina, San Diego, CA, USA). RNA was isolated from cells at three time-points (27, 40, and 48 h) during culturing at 28 °C, as described above. At each time-point, three samples were taken from different flasks. After washing cells with distilled water, total RNA was extracted following the manufacturer’s instructions (Invitrogen, Carlsbad, CA, USA) and purified using DNase and RNeasy micro kits (QIAGEN, Hilden, Germany). Transcriptome libraries for Illumina GAIIx were prepared with TruSeq Standard mRNA LT Sample Prep kit (Illumina, San Diego, CA, USA) and sequenced as per manufacturer’s instruction (paired-end sequences, 150 bp ×2). A total of 113.6 Gbp nucleotide (Gbn) sequence data were obtained, 71.8% of which (81.6 Gbn) had a quality value ≥30. These high-quality sequences were assembled with the Velvet/Oases assembler with a hash length of 37 [23]. Trinity assemblies were also used [24]. As a result, we obtained 112,852 contigs (29.3 Mbp) over 100 bp with an N50 size of 1575 bp. Of these, 88,570 (78%) had a blat alignment to the assembled genome (with default settings). A total of 68,571 of the 88,570 contigs were found to have complete open reading frames (ORFs) from the start to stop codon of at least 450 bp. Of these putatively full-length RNA contigs, 66,342 (97.3%) had a blat alignment to the assembled scaffolds. These data were used to produce gene models and annotation.

2.5. Gene Modeling

Gene predictions (Aurantiochytrium KH105 Gene Model ver. 1) were performed using ab initio and complementary DNA (cDNA) sequence alignment prediction derived from Augustus software [25]. Assembled Aurantiochytrium transcriptome sequences were processed with PASA [26], and the 500 longest ORFs were used to train the prediction software. Assembled Illumina RNA sequencing (RNA-seq) data were aligned to the assembled genome and incorporated as an Augustus “hint” on a repeat-masked genome produced using RepeatMasker [27].

2.6. Genome Browser

A genome browser was established for the assembled sequences using the Generic Genome Browser (GBrowser) 2.17 [28].

2.7. Annotation and Identification of Aurantiochytrium Genes

Protein-coding genes in the Aurantiochytrium KH105 genome were surveyed as follows. (i) Nucleotide and amino acid sequences of well annotated genes of model organisms were used as queries for basic local alignment search tool (BLAST) searches including TBLASTN (expect value < 1 × 10−10) of the Aurantiochytrium KH105 genome; (ii) Pfam domain searches were performed to identify protein domains included in putative proteins from all gene models (Pfam-A.hmm, release 24.0; http://pfam.sanger.ac.uk) [29]. Aurantiochytrium genes with conserved domains were included in a phylogenetic analysis; (iii) In order not to miss divergent homologs specific to Labyrinthula lineages, we confirmed them with at least one gene hit using reciprocal BLAST searches. Divergent homologs were shown as outgroups of annotated genes. RNA-seq reads were mapped to the assembled genome [29]. MAKER2 was used to tune-up the annotation and modeling [30].

2.8. Molecular Phylogeny

Molecular phylogenic analyses of phytoene desaturase (CrtI), geranylgeranyl phytoene synthase (CrtB), and geranylgeranyl phytoene synthase (CrtY) domains of CrtIBY were carried out using ClustalW [31]. Their amino acid sequences were aligned using default parameters and the resulting datasets were used to visualize maximum-parsimony trees with Sea View [32].

2.9. Analyses of Gene Expression Profiles

Differential gene expression profiles were analyzed using RNA-seq data with TopHat and Cluffinks, according to the method of Trapnell et al. [33]. The expression level of each gene was calculated as fragments per kilobase of transcript per million mapped reads (FPKM) with Cufflinks [34].

2.10. Functional Assay of crtIBY

After cultivation of Aurantiochytrium sp. KH105 in GPY medium for 4 days, total genomic DNA was extracted from cells using the phenol-chloroform method. The crtIBY1 gene was obtained using polymerase chain reaction containing 0.15 μg of genomic DNA, 1 unit of KOD FX neo (Toyobo, Osaka, Japan), 0.4 mM of each deoxynucleotide (dNTP), and 0.3 μM of each oligo DNA primer (5’-CAGTGTGCTGGAATTATGGCGCGCAGGGCGTCG-3’ and 5’-GATATCTGCAGAATTTCAGGCA TTCTTGTACAGCGGGAGC-3’) (Fasmac, Kanagawa, Japan) in 50 μL with incubation at 94 °C for 2 min and then 30 cycles of 98 °C for 10 s, 70 °C for 30 s, and 68 °C for 2 min. The crtIBY1 gene was ligated to the EcoRI-digested vector pYES2 (Thermo Fisher Scientific, Waltham, MA, USA) using the In-fusion HD system (Clontech Laboratories, Mountain View, CA, USA). Stellar competent Escherichia coli cells (Clontech Laboratories) were transformed using a heat shock method. Resultant pYES2-crtIBY and pYES2 plasmids were introduced individually to Saccharomyces cerevisiae INVSc1 (Thermo Fisher Scientific) using the polyethylene glycol/lithium acetate method. Transformants complementing uracil requirement were isolated and cultivated in uracil-defective minimum medium (0.67% yeast nitrogen base without amino acids (Difco Laboratories, Detroit, MI, USA), 0.19% yeast synthetic drop out medium without uracil (Thermo Fisher Scientific), 4% d-raffinose (Thermo Fisher Scientific) at 28 °C for 6 h. The crtIBY1 gene was expressed by addition of galactose (Thermo Fisher Scientific) to a final concentration of 2%, followed by further incubation for 16 h. For carotenoid analysis, cells from a 50-mL culture were collected by centrifugation (3500 rpm, 10 min, 4 °C), washed with water, and freeze-dried. Dried cells were crushed with glass beads (Wako Pure Chemical, Osaka, Japan) in 1 mL of chloroform/methanol (2:1, v/v) using a beads crusher (µT-12, Taitec, Saitama, Japan) at 2500 rpm for 120 s. The extract was recovered in the chloroform phase by centrifugation (3500 rpm, 10 min, 4 °C) of the crushed cells and evaporated under nitrogen gas spray. Carotenoids were dissolved with acetone/methanol (7:3, v/v) and analyzed on a 1260 Infinity HPLC system (Agilent Technologies, Santa Clara, CA, USA) equipped with a carotenoid separation column (S-5, 4.6 × 250 mm; YMC, Kyoto, Japan) and a photo diode array (Agilent Technologies) at 1 mL/min with acetonitrile/dichloromethane/methanol (7:2:1, v/v). Liquid chromatography-mass spectrometric (LC-MS) analysis was performed using LTQ Orbitrap XL (Thermo Fisher Scientific) equipped with a liquid chromatography (LC) column as mentioned above. We did not detect any oxidative decomposition products or other modified compounds.

3. Results and Discussion

3.1. Whole Genome Sequencing

Roche 454 WGS paired-end sequencing (661 bp on average read-length) yielded 2.9 Gb of nucleotide sequence data (Supplementary Table S1). In addition, Illumina Miseq paired-end sequencing (249 bp on average read-length, ×2) yielded 1.4 Gb of nucleotides, and GAIIx mate-pair sequencing (148 bp on average read-length, 3-kb library) yielded a total of 4.6 Gb. In total, 8.9 Gb sequence data were obtained.
The first sequence assembly yielded 2038 contigs with an N50 = 71.2 kb, and a total contig length of 76.5 Mb (Table S1). Subsequent scaffolding decreased the number of scaffolds to 512, N50 = 366.9 kb, and the total length of scaffolds was 76.7 Mb. On the other hand, k-mer analyses estimated the genome size of Aurantiochytrium sp. strain KH105 to be approximately 97 Mb (ranging from 94.7 to 104.1 Mb) (Supplementary Figure S1). This suggests that approximately 21% of the genome is likely repetitive or unassembled sequences. The original 8.9 Gb sequence data corresponded to approximately a 91.8-fold coverage of the genome. GC content was estimated from the assembled genome sequences to be 54.4% (Table 1; Supplementary Figure S2). This value is comparable to those of genomes for the Thraustochytrids, Aurantiochytrium sp. FCC1311 (56.8%) [15] and Schizochytrium sp. CCTCC M209059 (56.6%) [12], although the percentage is higher in Aurantiochytrium sp. T66 (62.8%) [13] and Thraustochytrium sp. ATCC 26185 (63.0%) [14] (Table 1). The GC content is also comparable for the oomycetes, Phytophthora infestans (51.0%), Phytophthora sojoe (54.4%), and Phytophthora romarum (54.4%) [35].

3.2. Gene Modeling

Gene models were constructed by combining RNA-seq sequence data (113.6 Gb of paired-end reads from the Illumina GAIIx) (Table S2) and other queries. There were 19,756 complete gene models with both start and stop codons (Table 1). Approximately 80% of them (15,800) were supported by RNA-seq analyses. At this modeling stage, the average gene length was 13,061 bp. The number of exons per gene was 4.7, and the average exon length was 513 bp, suggesting that an average gene consists of exons 2411-bp in length. This means that, on average, each gene contains an intron of nearly 10,565 bp. Here, we compared the gene content of Aurantiochytrium sp. strain KH105 to those of four other Thraustochytrid species (Table 1). The genome size of Aurantiochytrium sp. strain KH105 was almost double that of the other Thraustochytrid species. This suggests that the Aurantiochytrium sp. strain KH105 is diploid (see below).
The Aurantiochytrium sp. KH105 genome browser is accessible at Genome Projects, OIST Marine Genomics Unit website (http://marinegenomics.oist.jp/aurantiochytrium_sp_kh105, Okinawa Institute of Science and Technology, Okinawa, Japan).
To determine the ploidy level of the genome, we analyzed a set of 248 core eukaryotic genes (CEGs), which are highly conserved and are present in low copy numbers in higher eukaryotes [37]. For example, the yeasts, Saccharomyces cerevisiae and Schizosaccharomyces pombe, both haploid strains, have an average of 1.10 and 1.11 orthologs per CEG in their genomes, respectively (Supplementary Table S3), indicating that they are haploid. A similar analysis of genomes of Phytophthora infestans [38], Phytophthora ramorum, and Phytophthora sojae [39] showed that their average numbers were 1.25, 1.18, and 1.22, respectively. In contrast, the number of Auranthiochytrium sp. KH105 was 2.24, almost double those of the yeasts and Phytophthora. In addition, 95.6% of the 248 CEGs in the Auranthiochytrium sp. KH105 genome have two or more orthologs. Given the properties of CEGs, it is highly like that Auranthiochytrium sp. KH105 is diploid.

3.3. Carotenogenic Genes

A survey of the Aurantiochytrium genome for enzymes involved in carotenoid biosynthesis, using sequences from other eukaryotes and prokaryotes as queries, revealed 16 and 7 enzymatic genes involved in pathways for glycolysis and mevalonate biosynthesis, respectively (Supplementary Table S4). In addition, we characterized three genes from the isopentenyl pyrophosphate (IPP) to geranylgeranyl pyrophosphate (GGPP) pathway, including geranyl diphosphate synthase (GDPS), farnesyl diphosphate synthase (FDPS), and geranylgeranyl diphosphate synthase (crtE), and five genes from GGPP to astaxanthin pathway, including phytoene synthase (crtB), phytoene desaturase (crtI), lycopene cyclase (crtY), β-carotene ketolase (crtO), and β-carotene hydroxylase (crtZ) (Figure 1a; Supplementary Table S4). All eight of these genes appear duplicated. That is, two copies of GDPS were found in scaffolds 00034 and 00061, those for FDPS in scaffolds 00048 and 00052, those for crtE in scaffolds 00012 and 00031, those for crtB, crtI, and crtY in 00221 and 00288, those for crtO in 00001 and 00270, and those for crtZ in 00075 and 00157 (Figure 1b). RNA-seq analysis also confirmed the duplication of these genes.

3.4. Structure and Function of crtIBY Gene

In addition, we found that crtB, crt I, and crtY were clustered in scaffolds 00221 (tentatively called crtIBY1) and 00288 (crtIBY2), and that the gene order in both clusters was I-B-Y (Figure 2). Both expanded approximately 3.8 kb in the scaffolds and were expected to encode proteins of 1,268 amino acids. Molecular phylogeny indicated an affinity of KH105 CrtI (Figures S3 and S4), CrtB (Supplementary Figures S5 and S6), and CrtY (Figures S7 and S8) to archaean homologs. Messenger RNAs (mRNAs) transcribed by crtIBY1 and crtIBY2 (for example, base positions 106 to 3807 of crtIBY1 (Figure 2) had an open reading frame (ORF) without stop codons. This suggests that the mRNAs are translated to produce single polypeptides of 1268 amino acids.
Therefore, not only are crtB, crtI, and crtY clustered in the genome, but they are also fused into a single gene. Although there are several reports that show a cluster of two genes encoding carotenoid biosynthetic enzymes, for example, carRA in Phycomyces blakesleeanus [39] and crtYB in Xanthophyllomyces dendrorhous [49], this finding of a three-gene fusion appears novel. Since the biosynthetic process from GGPP to β-carotene is usually regulated by four independent enzymatic reactions catalyzed by CrtB, CrtI, and CrtY (green lines in Figure 1a), this gene fusion suggests that Aurantiochytrium has managed to increase efficiency by performing all three reactions with a single, multifunctional enzyme (Figure 1a). In addition, we have been unable to detect lycopene, which is an intermediate product of the pathway when the process is accomplished by four separate enzymatic reactions.
To confirm the function of crtIBY, the gene was heterologously expressed in yeast Saccharomyces cerevisiae INVSc1, which possesses GGPP synthase (BTS1), but lacks carotenoid biosynthetic genes, crtB, crtI, and crtY. Liquid chromatography–mass spectrometric analysis of carotenoids produced by the crtIBY1-expressing yeast identified a newly generated compound as β-carotene, based upon its mass profile, which is similar to that of authentic β-carotene (Figure 3). This indicates that the crtIBY gene encodes β-carotene synthase. The fact that levels of β-carotene increased in BTS1-overexpressing cells suggests that GGPP is the substrate of CrtIBY.

3.5. Gene Expression Profile

Carotenoid biosynthesis of Aurantiochytrium sp. KH105 becomes evident after 27 h of culture at 28 °C (mid-log phase of culture; Figure 4a), and carotenoid production plateaus after 72 h (starvation phase). Amounts of β-carotenes, canthaxanthin, and astaxanthin, plateaued at ≈48 h (stationary phase), 72 h, and 96 h, respectively, after initiation of incubation. We thought that these changes might reflect enzymatic gene expression; therefore, we performed a large-scale, gene expression profile analysis of the seven genes involved in carotenoid synthesis. This analysis showed that the concentration of FDPS mRNA decreased after 27 h of incubation (Figure 4b), while the concentrations of crtE, crtIBY, crtO, and crtZ mRNAs steadily increased from 27–48 h of incubation.
Major lipid biosynthetic pathways in Aurantiochytrium include those for polyunsaturated fatty acids (DHA and DPA) from acetyl-CoA, squalene, and sterols from farnesyl pyrophosphate (FPP), and xanthophylls from FPP/GGPP (Supplementary Figure S9). As mentioned above, in strain KH105, increased expression activity of crtE was evident, but overall expression levels of crtIBY, crtO, and crtZ appeared lower compared to that of crtE. Therefore, a key enzymatic reaction to synthesize xanthophylls is the CrtIBY-mediated reaction to produce β-carotene from GGPP. Although future studies should explore regulatory mechanisms of crtIBY gene expression in relation to crtE expression, genetic manipulation of the expression of these genes seems to be a key event that has allowed Aurantiochytrium to produce carotenoids more efficiently.
Since differential gene expression profile analysis showed fewer changes in the concentration of mRNAs for crtIBY, crtO, and crtZ, we further compared mRNA levels of the three genes between growth phases. Expression levels of crtO and crtZ increased from 27 (mid-log phase) to 40 h (late-log to stationary phase), and from 27 to 48 h (starvation phase), respectively (Figure 4c). Up-regulation of crtO and crtZ expression appeared to correspond well to increased concentrations of canthaxanthin and astaxanthin (Figure 4a,c). This suggests that genetic changes that up-regulate crtO and crtZ expression do increase carotenoid production. However, since overall levels of crtO and crtZ expression appear high during log phase compared to starvation phase, this explanation may be too simplistic. In contrast, the expression level of crtIBY appears lower than those of other enzymatic genes. Therefore, a plausible way to increase carotenoid production is to induce higher crtIBY expression during log, stationary, and starvation phases.

3.6. Occurrence of the crtIBY Gene in other Thraustochytrid Species

As described above, the genome of Aurantiochytrium sp. strain KH105 contains a fused gene, crtIBY, and the encoded protein likely acts as a trifunctional β-carotene synthase. Because genomes of other Thraustochytrid species have been sequenced (Table 1), we examined whether crtIBY is specific to Aurantiochytrium sp. 105 or whether it is also a feature of other Thraustochytrids. As a query of KH105 crtI, crtB, or crtY nucleotide sequences, we BLASTed genome sequences of the other four species. We found that all four Thraustochytrids possess a protein very similar to KH105 CrtIBY (Figure 5). No stop codons were found in the genomic region that includes the sequences. Especially, amino acid sequences of CrtIBY were almost completely identical among Aurantiochytrium sp. strain KH105 and Aurantiochytrium sp. FCC1311 and Schizochytrium sp. CCTCC M209059 (Figure 5). On the other hand, in CrtI of Aurantiochytrium sp. T66 and Thraustochytrium sp. ATCC 26185, there were three insertions consisting of two, three, or four amino acids. In addition, there were many amino acid substitutions in these two species, compared with the other three (Figure 5). Therefore, it is highly likely that a fused gene crtIBY is common to all Thraustochytrid species, although its function should be examined in each.

4. Conclusions

The decoded genome of Aurantiochytrium sp. strain KH105 demonstrated at least two genomic alterations associated with efficient production of carotenoids, gene duplication, and gene fusion. The fusion gene, crtIBY, looks common in all Thraustochytrid species. In addition, a large number of differential expression profiles of genes involved in carotenoid biosynthesis suggest key genetic manipulations that lead to more efficient production of carotenoids in this strain.

Supplementary Materials

Supplementary materials can be found at https://0-www-mdpi-com.brum.beds.ac.uk/2073-4425/9/4/200/s1.

Acknowledgments

We wish to thank Steven Aird for his help in preparation of the manuscript and two anonymous reviewers for helpful comments. This work was supported by Okinawa Intellectual Cluster Program, OIST, and in part by JST-CREST, Japan.

Author Contributions

T.A. and N.S. conceived and designed experiments. H.I., K.H., and R.K. carried out genome and transcriptomic analyses. H.I., R.S., A.N., K.W., and T.A. performed functional analyses. T.A., H.I., and N.S. wrote the manuscript with supervision of all the authors.

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Moss, S.T. (Ed.) Biology and Phylogeny of the Labyrinthulales and Thraustochytriales. In The Biology of Marine Fungi; Cambridge University Press: Cambridge, UK, 1986; pp. 105–129. [Google Scholar]
  2. Cavalier-Smith, T.; Allsopp, M.T.E.P.; Chao, E.E. Thraustochytrids are chromists, not Fungi: 18S rRNA signatures of Heterokonta. Phil. Trans. R. Soc. Lond. B 1994, 346, 387–397. [Google Scholar] [CrossRef]
  3. Yokoyama, R.; Honda, D. Taxonomic rearrangement of the genus Schizochytrium sensu lato based on morphology, chemotaxonomic characteristics, and 18S rRNA gene phylogeny (Thraustochytriaceae, Labyrinthulomycetes): Emendation for Schizochytrium and erection of Aurantiochytrium and Oblongichytrium gen. nov. Mycoscience 2007, 48, 199–211. [Google Scholar]
  4. Kimura, H.; Fukuba, T.; Naganuma, T. Biomass of Thraustochytrid protoctists in coastal water. Mar. Ecol. Prog. Ser. 1999, 189, 27–33. [Google Scholar] [CrossRef]
  5. Naganuma, T.; Takasugi, H.; Kimura, H. Abundance of Thraustochytrids in coastal plankton. Mar. Ecol. Prog. Ser. 1998, 162, 105–110. [Google Scholar] [CrossRef]
  6. Gupta, A.; Barrow, C.J.; Puri, M. Omega-3 biotechnology: Thraustochytrids as a novel source of omega-3 oils. Biotechnol. Adv. 2012, 30, 1733–1745. [Google Scholar] [CrossRef] [PubMed]
  7. Lewis, T.E.; Nichols, P.D.; McMeekin, T.A. Sterol and squalene content of a docosahexaenoic-acid-producing Thraustochytrid: Influence of culture age, temperature, and dissolved oxygen. Mar. Biotechnol. 2001, 3, 439–447. [Google Scholar] [CrossRef] [PubMed]
  8. Yamaoka, Y.; Carmona, M.L.; Oota, S. Growth and carotenoid production of Thraustochytrium sp. CHN-1 cultured under superbright red and blue light-emitting diodes. Biosci. Biotechnol. Biochem. 2004, 68, 1594–1597. [Google Scholar] [CrossRef] [PubMed]
  9. Li, Q.; Chen, G.Q.; Fan, K.W.; Lu, F.P.; Aki, T.; Jiang, Y. Screening and characterization of squalene-producing Thraustochytrids from Hong Kong mangroves. J. Agric. Food. Chem. 2009, 57, 4267–4272. [Google Scholar] [CrossRef] [PubMed]
  10. Kaya, K.; Nakazawa, A.; Matsuura, H.; Honda, D.; Inouye, I.; Watanabe, M.M. Thraustochytrid Aurantiochytrium sp. 18W-13a accumulates high amounts of squalene. Biosci. Biotechnol. Biochem. 2011, 75, 2246–2248. [Google Scholar] [CrossRef] [PubMed]
  11. Aki, T.; Hachida, K.; Yoshinaga, M.; Katai, Y.; Yamasaki, T.; Kawamoto, S.; Kakizono, T.; Maoka, T.; Shigeta, S.; Suzuki, O.; et al. Thraustochytrid as a potential source of carotenoids. J. Am. Oil Chem. Soc. 2003, 80, 789–794. [Google Scholar] [CrossRef]
  12. Ji, X.J.; Mo, K.Q.; Ren, L.J.; Li, G.L.; Huang, J.Z.; Haung, H. Genome sequence of Schizochytrium sp. CCTCC M209059, an effective producer of docosahexaenoic acid-rich lipids. Genome Announc. 2015, 3, e00819. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, B.; Ertesvåg, H.; Aasen, J.M.; Vadstein, O.; Brautaset, T.; Heggeset, T.M.B. Draft genome sequence of the docosahexaenoic acid producing Thraustochytrid Aurantiochytrium sp. T66. Genom. Data 2016, 8, 115–116. [Google Scholar] [CrossRef] [PubMed]
  14. Zhao, X.; Dauenpen, M.; Qu, X.; Qi, X. Genomic analysis of genes involved in the biosynthesis of very long chain polyunsaturated fatty acids in Thraustochytrium sp. 26185. Lipids 2016, 51, 1065–1075. [Google Scholar] [CrossRef] [PubMed]
  15. Sediki, K.; Godart, F.; Cigliano, R.; Sanseverino, W.; Barakata, M.; Ortet, P.; Rébellié, F.; Maréchal, E.; Cagnac, O.; Amato, A. Sequencing, de novo assembly, and annotation of the complete genome of a new Thraustochytrid species, strain CCAP_4062/3. Genome Annunc. 2018, 6, e01335-17. [Google Scholar] [CrossRef] [PubMed]
  16. Sambrook, J.; Russell, D.W. Isolation and Quantification of DNA. In Molecular Cloning: A Laboratory Manual, 4th ed.; Cold Spring Harbor Lab. Press: New York, NY, USA, 2001; pp. 1–80. [Google Scholar]
  17. Margulies, M.; Egholm, M.; Altman, W.E.; Attiya, S.; Bader, J.S.; Bemben, L.A.; Berka, J.; Braverman, M.S.; Chen, Y.J.; Chen, Z.; et al. Genome sequencing in microfabricated high-density picolitre reactors. Nature 2005, 437, 376–380. [Google Scholar] [CrossRef] [PubMed]
  18. Bentley, D.R. Whole-genome re-sequencing. Curr. Opin. Genet. Dev. 2006, 16, 545–552. [Google Scholar] [CrossRef] [PubMed]
  19. Dayarian, A.; Michael, T.P.; Sengupta, A.M. SOPRA: Scaffolding algorithm for paired reads via statistical optimization. BMC Bioinform. 2010, 11, 345. [Google Scholar] [CrossRef] [PubMed]
  20. Boetzer, M.; Henkel, C.V.; Jansen, H.J.; Butler, D.; Pirovano, W. Scaffolding pre-assembled contigs using SSPACE. Bioinformatics 2011, 27, 578–579. [Google Scholar] [CrossRef] [PubMed]
  21. Li, R.; Fan, W.; Tian, G.; Zhu, H.; He, L.; Cai, J.; Huang, Q.; Cai, Q.; Li, B.; Bai, Y.; et al. The sequence and de novo assembly of the giant panda genome. Nature 2010, 463, 311–317. [Google Scholar] [CrossRef] [PubMed]
  22. Marçais, G.; Kingsford, C. A fast, lock-free approach for efficient parallel counting of occurrences of k-mers. Bioinformatics 2011, 27, 764–770. [Google Scholar] [CrossRef] [PubMed]
  23. Zerbino, D.R.; Birney, E. Velvet: Algorithms for de novo short read assembly using de Bruijn graphs. Genom. Res. 2008, 18, 821–829. [Google Scholar] [CrossRef] [PubMed]
  24. Grabherr, M.G.; Haas, B.J.; Yassour, M.; Levin, J.Z.; Thompson, D.A.; Amit, I.; Adiconis, X.; Fan, L.; Raychowdhury, R.; Zeng, Q.; et al. Full-length transcriptome assembly from RNA-Seq data without a reference genome. Nat. Biotechnol. 2011, 29, 644–652. [Google Scholar] [CrossRef] [PubMed]
  25. Larkin, M.A.; Blackshields, G.; Brown, N.P.; Chenna, R.; McGettigan, P.A.; McWilliam, H.; Valentin, F.; Wallace, I.M.; Wilm, A.; Lopez, R.; et al. Clustal W and Clustal X version 2.0. Bioinformatics 2007, 23, 2947–2948. [Google Scholar] [CrossRef] [PubMed]
  26. Haas, B.J.; Delcher, A.L.; Mount, S.M.; Wortman, J.R.; Smith, R.K., Jr.; Hannick, L.I.; Maiti, R.; Ronning, C.M.; Rusch, D.B.; Town, C.D.; et al. Improving the Arabidopsis genome annotation using maximal transcript alignment assemblies. Nucleic Acid Res. 2003, 31, 5654–5666. [Google Scholar] [CrossRef] [PubMed]
  27. Smit, A.F.A.; Hubley, R.; Green, P. RepeatMasker Open-3.0. 1996–2010. Available online: http://www.repeatmasker.org (accessed on 13 December 2011).
  28. Stein, L.D.; Mungall, C.; Shu, S.; Caudy, M.; Mangone, M.; Day, A.; Nickerson, E.; Stajich, J.E.; Harris, T.W.; Arva, A.; et al. The generic genome browser: A building block for a model organism system database. Genome Res. 2002, 12, 1599–1610. [Google Scholar] [CrossRef] [PubMed]
  29. Finn, R.D.; Mistry, J.; Tate, J.; Coggill, P.; Heger, A.; Pollington, J.E.; Gavin, O.L.; Gunasekaran, P.; Ceric, G.; Forslund, K.; et al. The Pfam protein families database. Nucleic Acid Res. 2010, 38, D211–D222. [Google Scholar] [CrossRef] [PubMed]
  30. Holt, C.; Yandell, M. MAKER2: An annotation pipeline and genome-database management tool for second-generation genome projects. BMC Bioinform. 2011, 12, 491. [Google Scholar] [CrossRef] [PubMed]
  31. Li, L.; Stoeckert, C.J., Jr.; Roos, D.S. OrthoMCL: Identification of ortholog groups for eukaryotic genomes. Genome Res. 2003, 13, 2178–2189. [Google Scholar] [CrossRef] [PubMed]
  32. Gouy, M.; Guindon, S.; Gascuel, O. SeaView version 4: A multiplatform graphical user interface for sequence alignment and phylogenetic tree building. Mol. Biol. Evol. 2010, 27, 221–224. [Google Scholar] [CrossRef] [PubMed]
  33. Trapnell, C.; Williams, B.A.; Pertea, G.; Mortazavi, A.; Kwan, G.; van Baren, M.J.; Salzberg, S.L.; Wold, B.J.; Pachter, L. Transcript assembly and quantification by RNA-Seq reveals unannotated transcripts and isoform switching during cell differentiation. Nat. Biotechnol. 2010, 28, 511–515. [Google Scholar] [CrossRef] [PubMed]
  34. Trapnell, C.; Roberts, A.; Goff, L.; Pertea, G.; Kim, D.; Kelley, D.R.; Pimentel, H.; Salzberg, S.L.; Rinn, J.L.; Pachter, L. Differential gene and transcript expression analysis of RNA-Seq experiments with TopHat and Cufflinks. Nat Protoc. 2012, 7, 562–578. [Google Scholar] [CrossRef] [PubMed]
  35. Haas, B.J.; Kamoun, S.; Zody, M.C.; Jiang, R.H.; Handsaker, R.E.; Cano, L.M.; Grabherr, M.; Kodira, C.D.; Raffaele, S.; Torto-Alalibo, T.; et al. Genome sequence and analysis of the Irish potato famine pathogen Phytophthora infestans. Nature 2009, 461, 393–398. [Google Scholar] [CrossRef] [PubMed]
  36. GenBank (Internet). Bethesda (MD): National Library of Medicine (US), National Center for Biotechnology Information (1982). Available online: https://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/nuccore/BEYU00000000.1 (accessed on 13 December 2011).
  37. Parra, G.; Bradnam, K.; Ning, Z.; Keane, T.; Korf, I. Assessing the gene space in draft genomes. Nucleic Acid Res. 2009, 37, 289–297. [Google Scholar] [CrossRef] [PubMed]
  38. Tyler, B.M.; Tripathy, S.; Zhang, X.; Dehal, P.; Jiang, R.H.; Aerts, A.; Arredondo, F.D.; Baxter, L.; Bensasson, D.; Beynon, J.L.; et al. Phytophthora genome sequences uncover evolutionary origins and mechanisms of pathogenesis. Science 2006, 313, 1261–1266. [Google Scholar] [CrossRef] [PubMed]
  39. Arrach, N.; Fernández-Martín, R.; Cerdá-Olmedo, E.; Avalos, J. A single gene for lycopene cyclase, phytoene synthase, and regulation of carotene biosynthesis in Phycomyces. Proc. Natl. Acad. Sci. USA 2001, 98, 1687–1692. [Google Scholar] [CrossRef] [PubMed]
  40. Cunningham, F.X.; Gantt, E. Genes and enzymes of carotenoid biosynthesis in plants. Ann. Rev. Plant Phys. Plant Mol. Biol. 1998, 49, 557–583. [Google Scholar] [CrossRef] [PubMed]
  41. Cunningham, F.X., Jr.; Lee, H.; Gantt, E. Carotenoid biosynthesis in the primitive red alga Cyanidioschyzon merolae. Eukaryot. Cell 2007, 6, 533–545. [Google Scholar] [CrossRef] [PubMed]
  42. Takaichi, S. Carotenoids in algae: Distributions, biosynthesis and functions. Mar. Drug 2011, 9, 1101–1118. [Google Scholar] [CrossRef] [PubMed]
  43. Lemoine, Y.; Schoefs, B. Secondary ketocarotenoid astaxanthin biosynthesis in algae: A multifunctional response to stress. Photosynth. Res. 2010, 106, 155–177. [Google Scholar] [CrossRef] [PubMed]
  44. Lichtenthaler, H.K. The 1-deoxy-d-xylulose-5-phosphate pathway of isoprenoid biosynthesis in plants. ann. review plant physiol. Plant Mol. Biol. 1999, 50, 47–65. [Google Scholar] [CrossRef]
  45. Zhu, C.; Bai, C.; Sanahuja, G.; Yuan, D.; Farré, G.; Naqvi, S.; Shi, L.; Capell, T.; Christou, P. The regulation of carotenoid pigmentation in flowers. Arch. Biochem. Biophys. 2010, 504, 132–141. [Google Scholar] [CrossRef] [PubMed]
  46. Misawa, N. Carotenoid β-ring hydroxylase and ketolase from marine bacteria-promiscuous enzymes for synthesizing functional xanthophylls. Mar. Drug 2011, 9, 757–771. [Google Scholar] [CrossRef] [PubMed]
  47. Klassen, J.L. Phylogenetic and evolutionary patterns in microbial carotenoid biosynthesis are revealed by comparative genomics. PLoS ONE 2010, 5, e11257. [Google Scholar] [CrossRef] [PubMed]
  48. Schmidt, I.; Schewe, H.; Gassel, S.; Jin, C.; Buckingham, J.; Hümbelin, M.; Sandmann, G.; Schrader, J. Biotechnological production of astaxanthin with Phaffia rhodozyma/Xanthophyllomyces dendrorhous. Appl. Microbiol. Biotechnol. 2011, 89, 555–571. [Google Scholar] [CrossRef] [PubMed]
  49. Verdoes, J.C.; Krubasik, K.P.; Sandmann, G.; van Ooyen, A.J. Isolation and functional characterization of a novel type of carotenoid biosynthetic gene from Xanthophyllomyces dendrorhous. Mol. Gen. Genet. 1999, 262, 453–461. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Genes encoding carotenoid biosynthetic enzymes of Aurantiochytrium sp. KH105. (a) The biosynthetic pathway for astaxanthin includes eight major enzymes, including geranyl diphosphate synthase (GDPS), farnesyl diphosphate synthase (FDPS), geranylgeranyl diphosphate synthase (CrtE), phytoene synthase (CrtB), phytoene desaturase (CrtI), lycopene cyclase (CrtY), β-carotene ketorase (CrtO), and β-carotene hydroxylase (CrtZ). Green lines indicate common pathways shared by many prokaryotes and eukaryotes, while blue lines indicate other pathways [40,41,42,43,44,45,46,47,48]. In the genome of Aurantiochytrium sp. KH105, three genes for crtB, crtI, and crtY are consecutively fused into a single, gene, crtIBY. The fused gene likely promotes the process from geranylgeranyl pyrophosphate (GGPP) to β-carotene; (b) Two copies of GDPS, FDPS, crtE, crtIBY, crtO, and crtZ were found in different scaffolds. The numbers and lengths (kb) of scaffolds are shown, and directions of the genes are indicated by arrows.
Figure 1. Genes encoding carotenoid biosynthetic enzymes of Aurantiochytrium sp. KH105. (a) The biosynthetic pathway for astaxanthin includes eight major enzymes, including geranyl diphosphate synthase (GDPS), farnesyl diphosphate synthase (FDPS), geranylgeranyl diphosphate synthase (CrtE), phytoene synthase (CrtB), phytoene desaturase (CrtI), lycopene cyclase (CrtY), β-carotene ketorase (CrtO), and β-carotene hydroxylase (CrtZ). Green lines indicate common pathways shared by many prokaryotes and eukaryotes, while blue lines indicate other pathways [40,41,42,43,44,45,46,47,48]. In the genome of Aurantiochytrium sp. KH105, three genes for crtB, crtI, and crtY are consecutively fused into a single, gene, crtIBY. The fused gene likely promotes the process from geranylgeranyl pyrophosphate (GGPP) to β-carotene; (b) Two copies of GDPS, FDPS, crtE, crtIBY, crtO, and crtZ were found in different scaffolds. The numbers and lengths (kb) of scaffolds are shown, and directions of the genes are indicated by arrows.
Genes 09 00200 g001
Figure 2. Composition of crtIBY1 (a) and crtIBY2 (b), and their amino acid alignments with genes of other organisms. Darker lines indicate higher amino acid similarity. GenBank/DDBJ accession numbers are indicated in Supplementary Table S5.
Figure 2. Composition of crtIBY1 (a) and crtIBY2 (b), and their amino acid alignments with genes of other organisms. Darker lines indicate higher amino acid similarity. GenBank/DDBJ accession numbers are indicated in Supplementary Table S5.
Genes 09 00200 g002
Figure 3. Liquid chromatography–mass spectrometric analyses of carotenoids generated by yeasts expressing crtIBY from Aurantiochytrium sp. KH105. (a) Total ion chromatograms of carotenoids extracted from yeast cells carrying pYES2 or pYES2-crtIBY and the β-carotene standard. An ion signal was selected within the molecular weight range of 537.44–537.45; (b) Mass spectrometry of carotenoids extracted from yeast cells carrying pYES2 or pYES2-crtIBY and the β-carotene standard at the retention times of 13.35, 13.21, and 13.22 min, respectively.
Figure 3. Liquid chromatography–mass spectrometric analyses of carotenoids generated by yeasts expressing crtIBY from Aurantiochytrium sp. KH105. (a) Total ion chromatograms of carotenoids extracted from yeast cells carrying pYES2 or pYES2-crtIBY and the β-carotene standard. An ion signal was selected within the molecular weight range of 537.44–537.45; (b) Mass spectrometry of carotenoids extracted from yeast cells carrying pYES2 or pYES2-crtIBY and the β-carotene standard at the retention times of 13.35, 13.21, and 13.22 min, respectively.
Genes 09 00200 g003
Figure 4. Differential expression profiles of genes for carotenoid synthetic enzymes of Aurantiochytrium sp. KH105. (a) Time course of total carotenoid synthesis (left scale) by cells incubated at 28 °C in relation to cell number (right scale; DCW, dry cell weight); (b) Differential gene expression profile analysis showing changes during cultivation time from 27 h (mid-log phase), 40 h (late-log to stationary phase), and 48 h (starvation phase) in the concentration of messenger RNAs (mRNAs) of FDPS, crtE, crtIBY1, crtO, and crtZ. Three independent measurements were carried out. Fragments per kilobase of transcript per million mapped reads (FPKM), fragments per kilobase of exon per million reads mapped; (c) Comparison of the concentration of mRNAs of genes for crtIBY1, crtO and crtZ, between 27/27 h, 40/27 h and 48/27 h, respectively.
Figure 4. Differential expression profiles of genes for carotenoid synthetic enzymes of Aurantiochytrium sp. KH105. (a) Time course of total carotenoid synthesis (left scale) by cells incubated at 28 °C in relation to cell number (right scale; DCW, dry cell weight); (b) Differential gene expression profile analysis showing changes during cultivation time from 27 h (mid-log phase), 40 h (late-log to stationary phase), and 48 h (starvation phase) in the concentration of messenger RNAs (mRNAs) of FDPS, crtE, crtIBY1, crtO, and crtZ. Three independent measurements were carried out. Fragments per kilobase of transcript per million mapped reads (FPKM), fragments per kilobase of exon per million reads mapped; (c) Comparison of the concentration of mRNAs of genes for crtIBY1, crtO and crtZ, between 27/27 h, 40/27 h and 48/27 h, respectively.
Genes 09 00200 g004
Figure 5. Alignment of amino acid sequences of possible trifunctional β-carotenase synthases, CrtIBY. Sequences of CrtIBY1 and CrtIBY2 of Aurantiochytrium sp. KH105 are compared with those of Aurantiochytrium sp. FCC1311, Schizochytrium sp. CCTCC M209059, Aurantiochytrium sp. T66, and Thraustochytrium sp. ATCC 26185. Blue indicates CrtI, purple, CrtB, and green, CrtY. Asterisks indicate identical amino acid residues; colons denote difference between the upper three and the lower two species; hyphens indicate missing residues.
Figure 5. Alignment of amino acid sequences of possible trifunctional β-carotenase synthases, CrtIBY. Sequences of CrtIBY1 and CrtIBY2 of Aurantiochytrium sp. KH105 are compared with those of Aurantiochytrium sp. FCC1311, Schizochytrium sp. CCTCC M209059, Aurantiochytrium sp. T66, and Thraustochytrium sp. ATCC 26185. Blue indicates CrtI, purple, CrtB, and green, CrtY. Asterisks indicate identical amino acid residues; colons denote difference between the upper three and the lower two species; hyphens indicate missing residues.
Genes 09 00200 g005
Table 1. Genome assembly and annotation statistics for Aurantiochytrium sp. KH105 and their comparison with those of other Thraustochytrid species.
Table 1. Genome assembly and annotation statistics for Aurantiochytrium sp. KH105 and their comparison with those of other Thraustochytrid species.
Aurantiochytrium sp. KH105 (1)Aurantiochytrium sp. T66 (2)Aurantiochytrium sp. FCC1311 (3)Schizochytrium sp. CCTCC M209059 (4)Thraustochytrium sp. ATCC 26185 (5)
Genome
Estimated genome size (Mb)9543393939
Coverage (fold)8303 405562
Number of scaffolds1810184722323222250 (10,764)
N50 scaffold length (kb)4631342.8236.6595.8239.0 (2139)
Total scaffold length (Mb)86.043.438.939.138.6 (18.1)
Number of contigs55776833450416088130 (10,768)
N50 contig length (kb)105.712,95222,47452(2139)
Total contig length (Mb)7838.338.738.318
G+C content (%)54.462.856.856.663.0 (56.0)
Genes
Number of gene models16,98811,68311,853914210,797
Fused CrtIBY+++++
(1) The present study is accessible at Genome Projects, OIST Marine Genomics Unit website (http://marinegenomics.oist.jp/aurantiochytrium_sp_kh105, Okinawa Institute of Science and Technology, Okinawa, Japan); (2) Liu et al. [13]; (3) Sediki et al. [15,36]; (4) Ji et al. [12]; (5) Zhao et al. [14]. The number in parenthesis comes from the data bank.

Share and Cite

MDPI and ACS Style

Iwasaka, H.; Koyanagi, R.; Satoh, R.; Nagano, A.; Watanabe, K.; Hisata, K.; Satoh, N.; Aki, T. A Possible Trifunctional β-Carotene Synthase Gene Identified in the Draft Genome of Aurantiochytrium sp. Strain KH105. Genes 2018, 9, 200. https://0-doi-org.brum.beds.ac.uk/10.3390/genes9040200

AMA Style

Iwasaka H, Koyanagi R, Satoh R, Nagano A, Watanabe K, Hisata K, Satoh N, Aki T. A Possible Trifunctional β-Carotene Synthase Gene Identified in the Draft Genome of Aurantiochytrium sp. Strain KH105. Genes. 2018; 9(4):200. https://0-doi-org.brum.beds.ac.uk/10.3390/genes9040200

Chicago/Turabian Style

Iwasaka, Hiroaki, Ryo Koyanagi, Ryota Satoh, Akiko Nagano, Kenshi Watanabe, Kanako Hisata, Noriyuki Satoh, and Tsunehiro Aki. 2018. "A Possible Trifunctional β-Carotene Synthase Gene Identified in the Draft Genome of Aurantiochytrium sp. Strain KH105" Genes 9, no. 4: 200. https://0-doi-org.brum.beds.ac.uk/10.3390/genes9040200

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