Spotlight Selection
Applied and Industrial Microbiology
Research Article
16 April 2024

Identification and enzymatic properties of arginine decarboxylase from Aspergillus oryzae

ABSTRACT

Aspergillus oryzae spores, when sprinkled onto steamed rice and allowed to propagate, are referred to as rice “koji.” Agmatine, a natural polyamine derived from arginine through the action of arginine decarboxylase (ADC), is abundantly produced by solid state-cultivated rice koji of A. oryzae RIB40 under low pH conditions, despite the apparent absence of ADC orthologs in its genome. Mass spectrometry imaging revealed that agmatine was accumulated inside rice koji at low pH conditions, where arginine was distributed. ADC activity was predominantly observed in substrate mycelia and minimally in aerial mycelia. Natural ADC was isolated from solid state-cultivated A. oryzae rice koji containing substrate mycelia, using ammonium sulfate fractionation, ion exchange, and gel-filtration chromatography. The purified protein was subjected to sodium dodecyl sulfate poly-acrylamide gel electrophoresis (SDS-PAGE), and the detected peptide band was digested for identification by liquid chromatography-tandem mass spectrometry (LC-MS/MS). The gene AO090102000327 of strain RIB40 was identified, previously annotated as phosphatidylserine decarboxylase (PSD), and encoded a 483-amino acid peptide. Recombinant protein encoded by AO090102000327 was expressed in Escherichia coli cells cultivated at 20°C, resulting in the detection of 49 kDa and 5 kDa peptides. The protein exhibited pyruvoyl-dependent decarboxylase activity, favoring arginine over ornithine and showing no activity with phosphatidylserine. The gene was designated Ao-adc1. Ao-ADC1 expression in rice koji at pH 4–6 was confirmed through western blotting using the anti-Ao-ADC1 serum. These findings indicate that Ao-adc1 encodes arginine decarboxylase involved in agmatine production.

IMPORTANCE

Gene AO090102000327 in A. oryzae RIB40, previously annotated as a PSD, falls into a distinct clade when examining the phylogenetic distribution of PSDs. Contrary to the initial PSD annotation, our analysis indicates that the protein encoded by AO090102000327 is expressed in the substrate mycelia area of solid state-cultivated A. oryzae rice koji and functions as an arginine decarboxylase (ADC). The clade to which Ao-ADC1 belongs includes three other Ao-ADC1 paralogs (AO090103000445, AO090701000800, and AO090701000802) that presumably encode ADC rather than PSDs. Regarding PSD, AO090012000733 and AO090005001124 were speculated to be nonmitochondrial and mitochondrial PSDs in A. oryzae RIB40, respectively.

INTRODUCTION

Polyamines, including putrescine, spermidine, and spermine, have emerged as compelling candidates for mitigating the decline in the quality of life associated with aging (1). These biogenic amines extend longevity and ameliorate age-related pathologies, such as impaired locomotor activity (2), cognitive dysfunction (3), and chronic inflammation (4), primarily through the induction of autophagy in model organisms (5, 6). Among the natural polyamines, agmatine, a decarboxylated derivative of arginine, holds significant promise as a substance for promoting human health as shown in Fig. 1 (7, 8). A wealth of preclinical investigations has illuminated agmatine’s multifaceted modulatory effects on various molecular targets, positioning it as a potential therapeutic agent and nutraceutical (9). Remarkably, agmatine exhibits the capacity to scavenge reactive oxygen species, thereby safeguarding mitochondrial integrity in brain cells (10). In aged rats, an excessive accumulation of nitric oxide (NO) resulting from elevated NO synthase expression in the hippocampus and frontal cortex undergoes conversion to the deleterious oxidant peroxynitrite. This oxidative stress provokes inflammation and tissue damage, ultimately leading to cognitive deficits (11). Intriguingly, supplementation with agmatine significantly ameliorates age-related memory and learning impairments in rats by inhibiting NO synthase activities (12). Agmatine also regulates polyamine biosynthesis and transport by inducing antizyme (13). Furthermore, a detailed analysis has revealed that long-term oral agmatine intake counteracts hormonal imbalances, such as insulin resistance, while promoting urea synthesis and suppressing weight gain (obesity) induced by a high-fat diet. These beneficial effects are closely associated with agmatine-induced increases in metabolic rate, attributed to the upregulation of uncoupling proteins, and enhanced fatty acid oxidation mediated through the activation of carnitine biosynthesis. Carnitine plays a pivotal role in facilitating the transfer of long-chain fatty acids into mitochondria for β-oxidation (14).
Fig 1
Fig 1 Polyamine synthesis pathway of Aspergillus oryzae. Speculated pathways are indicated by arrows. Unidentified ADC is denoted by a dotted arrow.
Fermented foods are well-known for their abundance of polyamines, which originate from food ingredients like soybean, and are synthesized by microorganisms involved in the fermentation process. Notably, sake, a traditional Japanese rice wine, contains higher levels of agmatine (15, 16), although agmatine is absent in rice itself, indicating that its formation is attributed to microorganisms participating in the fermentation process (15, 16). The production of Japanese rice wine involves a complex fermentation process known as multiple parallel fermentation (MPF). During MPF, rice undergoes simultaneous saccharification by the filamentous fungus A. oryzae and ethanol fermentation by Saccharomyces cerevisiae (17). A. oryzae plays a vital role in the process, as it produces and secretes abundant hydrolyzing enzymes that effectively break down solid raw materials containing starches and proteins. As MPF progresses, nitric acid-reducing bacteria and lactic acid bacteria (LAB) spontaneously proliferate and produce nitrous acid and lactic acid, respectively, resulting in a reduction in environmental pH and safeguarding against contamination by undesirable microbes (17). In the previous study, it was shown that agmatine production was achieved in A. oryzae during solid-state cultivation, independent of S. cerevisiae. The fungus produced agmatine during solid-state cultivation, but not under submerged cultivation, and the productivity was substantially enhanced in response to acidic stimuli (18).
The enzyme arginine decarboxylase (ADC) (EC 4.1.1.19) catalyzes the decarboxylation of arginine, leading to the formation of agmatine (Fig. 1). However, genome analyses of A. oryzae suggest that it lacks orthologous genes encoding ADC in their genomes (19) and S. cerevisiae also lacks the orthologs (20). This observation aligns with previous research, indicating that fungi primarily synthesize putrescine from ornithine (Fig. 1) (21). Intriguingly, a recent study unveiled a unique agmatine catabolic pathway in Aspergillus niger involving a novel ureohydrolase called 4-guanidnobutyrase. This pathway converts agmatine to succinate via γ-aminobutyric acid (GABA), facilitated by a series of catabolic enzymes (Fig. 1) (22). Based on a comparative genomics analysis, A. oryzae may potentially harbor a corresponding agmatine catabolic pathway, where arginine is converted to putrescine via ornithine, serving as the principal route for polyamine biosynthesis (19). Efficient agmatine production by A. oryzae is attributed to the induction of an unidentified low-pH-dependent ADC and/or the low pH-dependent inhibition of agmatine catabolic pathway enzymes such as agmatine oxidase (amine oxidase) and/or agmatinase (agmatine amidinohydrolase) during solid-state cultivation (Fig. 1).
The structure of mammalian ADC was recently reported, showing distinct characteristics from bacterial and plant ADCs but sharing similarities with mammalian ornithine decarboxylase (ODC) (23). A. oryzae is known to possess four ODCs (ODC1–ODC4). However, recombinant ODC1 and ODC2 exhibited no ADC activity at acidic pH (pH 4.0), indicating the possible involvement of other decarboxylases or an unidentified ADC in agmatine production (18). In the present study, our aim is to purify natural ADC from A. oryzae hyphae and identify this unidentified ADC.

RESULTS

Growth dependency of synthesis of arginine and downstream metabolites in rice koji

The growth dependency of the synthesis of arginine and downstream metabolites in rice koji was examined by mass spectrometry imaging (MSI) (24). We focused on arginine, agmatine, putrescine, spermidine, 4-guanidinobutyric acid, and γ-aminobutyric acid (GABA), for which ionization conditions have been established, allowing identification based on m/z values. As shown in Fig. 2, these compounds were not detected in steamed rice; however, as A. oryzae grew, their accumulation was observed with increasing incubation time. Arginine and spermidine showed accumulation around the edges in the rice koji cultivated for 19 and 27 hours but were detected inside the rice koji cultivated for 45 hours. At 45 hours, where arginine was strongly detected in the rice koji, accumulation of polyamines (putrescine and spermidine) was observed. Agmatine was detected slightly in the peripheral region in the rice koji section. These results suggested that arginine and downstream metabolites are synthesized in response to mycelial growth.
Fig 2
Fig 2 Growth dependency of the synthesis of arginine and downstream metabolites in steamed rice and rice koji. Metabolites include arginine (m/z 175.12), agmatine (m/z 131.13), putrescine (m/z 89.11), spermidine (m/z 146.16), 4-guanidino butyric acid (m/z 146.09), and GABA (m/z 104.07). Accompanying the mass spectrometry imaging results are optical images of steamed rice and rice koji sections. The scale bar is 1 mm.

pH dependency of agmatine localization in rice koji

In the previous study, the amount of agmatine in rice koji increased when cultivation pH was lowered to 3 (18). To examine the localization of arginine and agmatine, imaging patterns were compared. As shown in Fig. 3; Fig. S1, arginine was detected more strongly in rice koji at pH 3.0 and 6.0 than in steamed rice. On the other hand, agmatine localization was not observed in steamed rice and rice koji at pH 6.0, consistent with previous data (18). Furthermore, the accumulation of agmatine overlapped with parts of the region where arginine accumulation was observed. This result suggests that agmatine decarboxylase, the enzyme responsible for producing agmatine from arginine, is also produced in the interior of the steamed rice, likely by the nutrient mycelia.
Fig 3
Fig 3 pH dependency of arginine (m/z 175.12) and agmatine (m/z 131.13) localizations in steamed rice and rice koji. Optical images of steamed rice and rice koji sections accompany the mass spectrometry imaging results. The scale bar is 500 µm.

Comparison of ADC activity in aerial and substrate mycelia

As agmatine was detected in the interior region of rice koji by MSI analysis, ADC was expected to exist in the mycelium. To compare ADC activity in aerial and substrate mycelia, crude cell extract was obtained from each fraction. A. oryzae RIB40 conidia were inoculated into steamed rice and mixed well. Solid-phase cultivation was conducted to obtain rice koji. The obtained rice koji was suspended and vortexed vigorously in phosphate buffer and centrifuged (7,000 × g). By repeating those steps, aerial and substrate mycelial fractions were separated as supernatants and precipitates (Ps), respectively, as described in the Materials and Methods section. The collected precipitates were used for the substrate mycelial fraction. Aerial mycelium in the supernatant was collected by centrifuging at a higher speed (10,000 × g). Substrate and aerial fractions were frozen with liquid nitrogen and disrupted by a mixer and sonication on ice. The obtained mycelial and aerial mycelial fractions of rice koji were used for the ADC assay. As shown in Table 1, 83% of the total mycelium was obtained as substrate mycelia. Most of the ADC activity (98.4%) was detected in substrate mycelia. When each activity was compared per hyphal amount, 92.6% was also found in substrate mycelium. Based on the ratio, ADC was mainly expressed in substrate mycelia.
TABLE 1
TABLE 1 ADC activity in substrate and aerial mycelia fractions in A. oryzae RIB40
FractionADC activity
[(nmolagmatine) min−1 (gpowdered steamed rice)−1]
GlcNAc
(ng/gculture)
Substrate mycelia4.64 × 10266.3
Aerial mycelia7.5413.5

Isolation and purification of natural ADC from A. oryzae RIB40

Since ADC activity was mainly detected in the substrate mycelia, purification of the enzyme was attempted using these mycelia. The substrate mycelia fraction obtained from 300 g of rice koji was frozen with liquid nitrogen, disrupted using a mixer, and sonicated on ice. The disrupted mycelial sample was centrifuged, and the supernatant was subjected to ammonium sulfate fractionation. The resulting active fraction was dialyzed and applied to an anion-exchange column, eluted with a linear gradient of NaCl. The fractions containing ADC activity were collected, concentrated, and further applied to a gel filtration column. Active fractions were concentrated and loaded onto an SDS-PAGE as shown in Fig. 4A. In gel-filtration chromatography, ADC activity was detected in several fractions in Fig. 4B. Fraction 14 exhibited the highest purification fold based on the ADC-specific activity (0.258 pmolagmatine/min/mgprotein) in Fig. 4B.
Fig 4
Fig 4 SDS-PAGE demonstration of purified ADC after gel-filtration chromatography. (A) SDS-PAGE analysis of eluted fractions after gel-filtration chromatography. The band of interest is indicated by a black arrow. (B) Molecular mass estimation of ADC. Molecular mass standards used were: low-molecular weight (LMW) marker, ovalbumin (44,000 Da), conalbumin (75,000 Da), aldolase (158,000 Da), ferritin (440,000 Da), and thyroglobulin (669,000 Da). Left, molecular mass [Da]; right, ADC activity [(pmolagmatine) min−1 (mgprotein)−1]. Squares, molecular mass; closed circles, ADC activity. The dotted line indicates the highest active fraction.

Protein identification and its phylogenetic analysis

The active fractions obtained after gel filtration were concentrated and applied to an SDS-PAGE preparatory gel. Two thick bands of fraction 14 were excised, dehydrated in acetonitrile, and alkylated. The resulting product was then digested with trypsin and desalted. The isolated peptides were analyzed using liquid chromatography-tandem mass spectrometry (LC-MS/MS). The effluent from the high performance liquid chromatography (HPLC) was directly electrosprayed into the mass spectrometer, where peptides separated by LC were ionized. Experimental conditions are described in Table S1. MS1 survey scans were acquired, and peptides were identified. MS/MS fragmentation spectra were searched with Proteome Discoverer 2.4 (Thermo Fisher Scientific K.K.) against the A. oryzae RIB40 genome database. The results were further analyzed using Mascot™ v. 2.6 (Matrix Science, SC), and the identified proteins are listed in Table S2. Among them, gene AO090102000327 (A. oryzae RIB40 chromosome 4) encoding a 54 kDa protein showed the highest mascot score (895), and it was annotated as phosphatidylserine decarboxylase (PSD). In the A. oryzae genome database, no ADC was annotated, and six PSDs (AO090012000733, AO090005001124, AO090103000445, AO090102000327, AO090701000800, and AO090701000802) were annotated. To determine the phylogenetic relationships of PSDs in A. oryzae RIB40, PSDs from various organisms were selected, and their amino acid sequences were compared. Based on the aligned data, a phylogenetic tree was constructed as shown in Fig. 5. PSDs were classified into three major clades: the first one is the non-mitochondria-derived group, containing one gene (AO090012000733) in A. oryzae RIB40; the second one is the mitochondria-derived group, including bacteria, (2527) with one gene (AO090005001124) in A. oryzae RIB40; and the third one is a unique clade, with no enzymatic activity reported for PSDs and containing four genes (AO090103000445, AO090102000327, AO090701000800, and AO090701000802) in A. oryzae RIB40 (28, 29). The protein identified in the present study (Gene AO090102000327) belongs to the third clade. PSD has been reported to function as a pyruvoyl-dependent decarboxylase, and unlike many decarboxylases, it does not require pyridoxal phosphate (PLP) (30, 31) as a cofactor. The inactive proenzyme of the pyruvoyl-dependent enzyme is cleaved into two active subunits, generally designated as α and β subunits. The cleavage occurs at the serine residue, termed nonhydrolytic serinolysis, and the NH2 terminal of the serine residue is converted to ammonia and the pyruvoyl group (32). Gene AO090102000327, which was predicted to encode ADC in this study, had no introns in its open reading frame. Speculated amino acid sequence encoded by AO090102000327 possesses the Gly-Gly-Ser-Thr region that contains Ser functional for serinolysis of pyruvoyl-dependent enzyme (3336), resulting in 441 amino acid peptide (49 kDa) and 42 amino acid one (5 kDa). To further elucidate the function of the enzyme encoded by AO090102000327, we attempted to express it as a recombinant form in Escherichia coli.
Fig 5
Fig 5 Phylogenetic tree based on amino acid sequences of phosphatidylserine decarboxylase homologs. This tree was constructed using the randomized axelerated maximum likelihood (RAxML) method provided by GenomeNet and operated by the Kyoto University Bioinformatics Center. Bootstrap resampling was performed 1,000 times, and only values observed in more than 50% of the replicas are shown at the branching points. The scale bar indicates 1 substitution per one amino acid. This phylogenetic tree includes six PSD orthologs of A. oryzae RIB40 (AO090012000733, AO090005001124, AO090103000445, AO090102000327, AO090701000800, and AO090701000802). The arrows indicate ADC and PSD genes in A. oryzae RIB40. The ADC1 identified in the present study (Gene AO090102000327) is indicated by a thick arrow.

Enzymatic characteristics of recombinant protein encoded by AO090102000327

The full-length open reading frame of AO090102000327 was amplified via polymerase chain reaction (PCR) and introduced into the expression plasmid pET28a to enable the expression of the encoded protein in recombinant form. The resulting plasmid, designated pET28a-Ao-ADC1, was then introduced into E. coli BL21 CodonPlus (DE3)-RIL cells. Expression of the encoded protein with N-terminal His tags was induced by the addition of isopropyl β-D-1-thiogalactopyranoside (IPTG). Following induction, cells were harvested through centrifugation, resuspended in phosphate buffer, and disrupted using sonication. After the removal of cell debris via centrifugation, the supernatant and precipitate were combined with an SDS-containing sample buffer and analyzed using SDS-PAGE. Upon incubation of cells at 37℃, the induced protein was detected in insoluble precipitates as a 54 kDa protein (Fig. 6A). To mitigate protein aggregation, E. coli cells were cultivated at 20℃ following IPTG induction. The induced proteins were detected in the soluble fraction with sizes of 54 kDa and 49 kDa. The 49 kDa size aligned well with the anticipated size, resulting from the serinolysis of the pyruvoyl-dependent enzyme. However, the smaller fragment (5 kDa) was not detected in the SDS-PAGE, possibly due to being beyond the separation range. Protein purification was carried out using crude extracts obtained from cells cultivated at 20°C. After removing cell debris by centrifugation, the supernatant was applied to a column packed with Ni resin. Elution was performed stepwise with phosphate buffer supplemented with imidazole, followed by dialysis against phosphate buffer as described in the Materials and Methods section. The protein was purified to homogeneity, as shown in Fig. 6B. However, the 5 kDa subunit could not be confirmed in the normal SDS-PAGE. Therefore, Tris-Tricine SDS-PAGE was performed, and proteins were visualized by silver staining as shown in Fig. 6C. The 5 kD band corresponding to α subunit (shown as MA) was detected with 45 kD β subunit (shown as MB). To examine the enzymatic characteristics of the purified protein, the decarboxylation activity of arginine, ornithine, and phosphatidylserine was assessed, along with kinetic parameters (Km and Vmax). Reaction mixtures containing the substrate (arginine, ornithine, or phosphatidylserine) were incubated at various temperatures (20, 30, 40, 50, or 60℃) for different reaction times (5–60 min) with the purified protein in citrate buffer at various pH values (pH 3.0, 4.0, 5.0, or 6.0). The reactions were terminated by the addition of tricarboxylic acid (TCA), and the reaction mixtures were filtered. Contents of agmatine, putrescine, and phosphatidylethanolamine were determined to assess ADC, ODC, and PSD activities, respectively. Kinetic parameters (Km and Vmax) were then determined for arginine, ornithine, and phosphatidylserine. As shown in Table 2, the protein exhibited the lowest Michaelis constant (25 µM) and the highest maximum velocity (5.0 × 10−1 mmol/min) when arginine was used as a substrate. On the other hand, no PSD activity was detected. Based on the obtained values, we concluded that AO090102000327 of A. oryzae RIB40 encodes arginine decarboxylase, not phosphatidylserine decarboxylase, and designated AO090102000327 as Ao-adc1. As shown in Table 3, no significant difference in decarboxylation activity was observed between the presence and absence of PLP, indicating that the protein is a pyruvoyl-dependent decarboxylase and does not require PLP for its activity. The optimal pH and temperature for ADC activity were also evaluated as shown in Fig. 7, showing that Ao-ADC1 optimally functions at pH 4.0 and 40℃.
Fig 6
Fig 6 SDS-PAGE demonstration of expressed recombinant Ao-ADC1 in E. coli and purified Ao-ADC1. (A) E. coli BL21(DE3) codon plus cells harboring pET28a or pET28a-ADC1 were cultivated at 37°C, and Ao-ADC1 expression was induced IPTG addition, followed by further cultivation for 4 hours at 37°C or 20°C. +, induction; −, no induction. Proteins in SDS-PAGE were visualized by Coomassie brilliant blue staining. (B) SDS-PAGE demonstration of purified Ao-ADC1 visualized by Coomassie brilliant blue staining. (C) Tris-Tricine SDS-PAGE demonstration of purified Ao-ADC1 visualized by silver staining. M, molecular mass markers; C, crude extract obtained from E. coli cells cultivated at 20℃ with IPTG addition; N, purified Ao-ADC1. Black arrows, Precursor of Ao-ADC1; White arrow with MA, processed α subunit of Ao-ADC1; White arrow with MB, processed β subunit of Ao-ADC1.
Fig 7
Fig 7 Enzyme characteristics of Ao-ADC1. ADC activity [(pmolagmatine) min−1] of purified Ao-ADC1 was measured at selected pH values (3.0, 4.0, 5.0, or 6.0) and temperatures (20, 30, 40, or 50°C). (A) Effect of temperature on Ao-ADC1 activity. (B) Effect of pH on Ao-ADC1 activity.
TABLE 2
TABLE 2 Kinetic parameters of recombinant Ao-ADC1 for various substrates
SubstrateKm
(mM)
Vmax
(μmol/min)
Arginine2.52 × 10−25.01 × 10−1
Ornithine70.71.43 × 10−7
PhosphatidylserineN.D.aN.D.a
a
Not detected.
TABLE 3
TABLE 3 ADC activity of recombinant Ao-ADC1 and crude extract of E. coli cells in the presence (+) and absence (−) of PLP
PLPADC activity (nmol/min/mg)
Recombinant Ao-ADC1Crude extract of E. coli cells
1.15 ± 0.14N.D.a
+1.22 ± 0.2073.3 × 10−3
a
Not detected.

Effect of pH on ADC1 expression in A. oryzae RIB40

The fungus produces agmatine during solid-state cultivation but not under submerged cultivation, and its productivity is notably enhanced in response to acidic conditions (18). To investigate the effect of pH on ADC1 expression in A. oryzae RIB40, we employed immunodetection using specific polyclonal antisera. A. oryzae RIB40 was cultured for 45 hours at 30°C under solid-state conditions with a pH range of 3.0–6.0. Disrupted crude cell extracts containing mycelia were obtained and subjected to SDS-PAGE, followed by electroblotting onto polyvinylidene difluoride (PVDF) membranes. Western blotting analysis was performed using these antibodies, and immunocomplexes were detected as shown in Fig. 8. ADC1 expression was detected under cultivation conditions with pH ranging from 4 to 6. However, at pH 3, ADC1 expression was notably lower.
Fig 8
Fig 8 Effect of incubation pH on ADC1 expression in rice koji. A. oryzae RIB40 was cultured for 45 hours at 30°C under solid-state conditions with a pH range of 3.0–6.0. Disrupted crude cell extracts containing mycelia were obtained and subjected to SDS-PAGE, followed by electroblotting onto PVDF membranes. Rabbit anti-Ao-ADC1 serum was used to detect Ao-ADC1, and immunocomplexes with Alexa Fluor 700 goat anti-rabbit IgG were visualized with an Odyssey infrared imaging system. (A) SDS-PAGE demonstration with Coomassie brilliant blue staining. (B) Western blotting patterns on PVDF membrane. M, molecular mass standards; Arrow, position of Ao-ADC1.

DISCUSSION

Although the growth-dependent accumulation of amino acids inside rice koji has been previously examined (37), the localizations of arginine and downstream metabolites in rice koji have not been reported. The accumulation of polyamines (putrescine and spermidine) observed in rice koji, as depicted in Fig. 2, along with the significant detection of arginine, suggests that these metabolites are synthesized from arginine and accumulate during the growth of A. oryzae. Moreover, the accumulation of agmatine was visualized in frozen sections of rice koji at pH 3.0 (Fig. 3). This result is consistent with a previous study that agmatine is highly accumulated in low pH environments (18). Interestingly, this accumulation of agmatine overlapped with the region where arginine accumulation was observed. This result suggests that the A. oryzae-derived ADC is presumably present in the substrate mycelia of A. oryzae. Indeed, when ADC activity was examined in aerial and substrate mycelia, high activity was observed in substrate mycelia (Table 1). However, A. oryzae lacks orthologous genes encoding ADC in their genomes (19); hence, we attempted to purify ADC from these substrate mycelia of A. oryzae. As a result of purification and identification of target protein, the gene AO090102000327 was identified as shown in Table S2, which was previously annotated as encoding PSD (3840) in the database. By comparing the selected amino acid sequences of PSDs from various organisms, a phylogenetic tree was constructed as shown in Fig. 5. The protein identified in the present study (Gene AO090102000327) belonged to the clade with no enzymatic activity reported for PSDs. This suggests that the PSDs in this clade are a group of enzymes with ADC activity.
The recombinant protein product, derived from AO090102000327, exhibited agmatine decarboxylase (ADC) activity rather than PSD activity as shown in Table 2. This discrepancy is not uncommon, as the nomenclature of decarboxylases is often based on amino acid sequence similarity. A similar misidentification has been seen with human ADC, previously identified as ODC. By the thorough enzyme characterization, the recombinant protein that previously identified ODC was indeed ADC (23). We identified the protein encoded by this gene as ADC and named it Ao-ADC1. This is the first report of the discovery of ADC in filamentous fungi. Although molecular mass of Ao-ADC1 from A. oryzae was 45 kDa as shown in Fig. 4, molecular mass of recombinant Ao-ADC1 in E. coli was 54 kDa from the database in the LC-MS/MS result (Table S2). This is presumably due to the self-cleavage accompanying enzyme maturation. When E. coli cells were incubated at 37℃, the induced protein was detected in insoluble precipitates as a 54 kDa protein. However, when E. coli cells were induced at 20℃ by IPTG to prevent protein aggregation, the induced proteins were detected in the soluble fraction with sizes of 54 kDa and 49 kDa (Fig. 6A). Furthermore, when the enzyme expressed in E. coli was purified, only 49 kDa protein fragment was confirmed, and 5 kDa protein fragment was detected using Tris-Tricine SDS-PAGE, as shown in Fig. 6B and C. The 49 kDa size aligned well with the anticipated size resulting from the serinolysis of the pyruvoyl-dependent enzyme (41). Therefore, it is likely that 5 and 49 kDa fragments form a heterocomplex as β-subunit and α-subunit of Ao-ADC, respectively, and likely function as a pyruvoyl-dependent decarboxylase. Moreover, functional analysis of E. coli ADC reveals PLP-dependent activity (21). However, as shown in Table 3, the addition of PLP did not enhance enzymatic activity in Ao-ADC1, indicating that this enzyme functions independently of PLP. In order to gain precise insights into the reaction mechanisms of Ao-ADC1, analysis such as protein crystallization of Ao-ADC1 is needed. Furthermore, the molecular mass of the purified enzyme estimated by gel filtration was higher than 440 kDa, whereas the size estimated by SDS-PAGE was about 45 kDa (Fig. 4A), suggesting that natural ADC possesses an oligomeric form.
As shown in Fig. 3, agmatine accumulates more prominently in steamed rice at pH 3 compared with that at pH 6. These findings suggest that the biosynthesis of agmatine may be responsive to acid stress. In E. coli, it has been documented that the acid-tolerant mechanism involves the consumption of protons through the decarboxylation reaction of arginine to agmatine (39, 40). A. oryzae may employ a similar acid tolerance mechanism through the biosynthesis of agmatine. However, as shown in Fig. 8, the expression levels of Ao-ADC1 in A. oryzae at pH 3 were lower than those at pH 4–6. The significant agmatine accumulation at pH 3 cannot be solely attributed to the increased expression of Ao-ADC1. It is hypothesized that the expression and activity of agmatine oxidase (amine oxidase) and/or agmatinase (agmatine amidinohydrolase) may be suppressed at pH 3. To substantiate this hypothesis, it is necessary to conduct gene knockout experiments targeting these genes and subsequently assess agmatine accumulation levels.

MATERIALS AND METHODS

Microbial strains and plasmid

A. oryzae RIB40 (19) was obtained from the National Research Institute of Brewing (Higashihiroshima, Hiroshima, Japan). E. coli strain TG1 (Funakoshi Co., Ltd., Tokyo, Japan) (Table S3), which was used for DNA manipulation to construct the plasmid pET28a derivatives to express Ao-adc1 gene from A. oryzae RIB40 (19), was routinely cultivated at 37°C in the lysogeny broth (LB) medium containing 20 µg/mL kanamycin. For the Ao-ADC1 production, E. coli BL21 Codon Plus (DE3)-RIL (Agilent, Santa Clara, CA, USA) (Table S3) cells harboring the expression plasmid were cultivated at 37°C or 20℃ on LB medium containing 20 µg/mL kanamycin and 30 µg/mL chloramphenicol. Plasmid pET28a was purchased from Merck (Darmstadt, Germany) and used for recombinant ADC1 expression in E. coli BL21 Codon Plus (DE3)-RIL cells.

Preparation of a solid culture (rice koji) of A. oryzae

Rice koji was prepared as previously described (37). Rice grains of 300 g polished to 90% of the total weight were soaked in tap water for 3 hours and steamed for 50 min. After cooling to 30°C, 3 g of the A. oryzae RIB40 conidia were inoculated into steamed rice and mixed well. The solid phase medium was cultivated with water-moistened gauze (15 × 23 cm), wrapped in a cotton cloth (35 × 35 cm) in a plastic container (16 × 23 × 5 cm), and incubated at 30°C. After 19 and 27 hours, the rice grains containing A. oryzae were mixed upside down to improve the aeration and lower the temperature. Finally, the solid culture was collected as rice koji after 45 hours cultivation and cooled at room temperature for 2 hours. The obtained rice koji was used in subsequent experiments.

Rice koji sectioning and mass spectrometry imaging analysis

Sample preparation was carried out according to the reported procedure (42) with slight modifications. One rice koji or steamed rice was first embedded in a mold (Base mold A, Falma, Tokyo, Japan) with 10% gelatin solution and then frozen with liquid nitrogen. The frozen sample block was attached to the sample holder using the optimal cutting temperature compounds (Surgipath FSC 22, Leica Biosystems, Nussloch, Germany) before sectioning using a cryostat (CM1850, Leica Biosystems). Longitudinal sections were obtained with 14 µm thickness from the approximate center of the solid phase medium grain. Adhesive cryofilm (Section-Lab, Yokohama, Kanagawa, Japan) was used to acquire the sample sections. Finally, the sections were attached to indium tin oxide (ITO) glass (Matsunami Glass, Osaka, Japan) using conductive double-sided tape (Shielding Non-woven Fabric Tape; 3M Company St. Paul, MN, USA). The matrix, α-cyano-4-hydroxycinnamic acid (CHCA) (Merck, Darmstadt, Germany), was supplied to the surface of the sections at 220°C with a thickness of 0.7 µm using a vacuum sublimation machine, iMLayer (Shimadzu, Kyoto, Japan). After that, 500 µL of 90% CHCA was sprayed evenly onto the glass slide using an airbrush. Then, the analysis was performed using iMScope QT (Shimadzu). All mass spectra were obtained in the m/z range (88.00–210.00) in positive ion mode. After analysis, analysis software IMAGEREVEAL MS (Shimadzu) was used to create the MS images by selecting m/z derived from target metabolites.

Separation of aerial and nutrient mycelium fraction

When aerial and substrate mycelial fractions were needed, obtained rice koji was suspended in 10 mL of 50 mM phosphate buffer (pH 7.0) in a 50 mL centrifuge tube and vortexed vigorously. The suspended solution was centrifugation at 7,000 × g, 4°C, for 10 min, and the supernatant and precipitate were separately collected. To the precipitate, 10 mL of 50 mM phosphate buffer (pH 7.0) was added again, vortexed vigorously, and centrifuged at 7,000 × g, 4°C, for 10 min. This step was repeated six times, and the final precipitate was frozen with liquid nitrogen, and the supernatants (S) (60 mL in total) were collected and centrifuged at 10,000 × g, 4°C for 10 min to collect the aerial mycelium as a precipitate. The precipitate (aerial mycelial fraction) was then frozen with liquid nitrogen and resuspended in 70 mL of 50 mM phosphate buffer (pH 7.0), followed by crushing with mixer and sonication on ice. The mixture was used as an aerial mycelial fraction. The precipitate containing substrate mycelia obtained by the centrifugation was frozen with liquid nitrogen. To 10 g of the precipitate, 100 mL of 50 mM phosphate buffer (pH 7.0) was added, followed by crushing with mixer and sonication on ice. The mixture was used as a substrate mycelial fraction.

Monitoring of the hyphal growth of A. oryzae

The growth of A. oryzae was evaluated by determining the amount of N-acetylglucosamine (GlcNAc), which is the building block of the major fungal cell wall constituent, chitin (18). To measure the total amount of mycelia, 2 g of the rice koji was dried at 100°C for 1 hour and completely ground using a mortar. The resultant powder was suspended in 10 mL of 50 mM phosphate buffer (pH 7.0), vigorously vortexed for 10 seconds, and recovered by centrifugation (10,000 × g, 10 min). These washing steps were repeated three times, and the resultant pellet was resuspended in 10 mL 50 mM phosphate buffer (pH 7.0). Next, 10 mg of Yatalase (TaKaRa Bio, Tokyo, Japan) was added to the suspension and incubated at 37°C for 1 hour, with reciprocal shaking at 200 rpm. The supernatant was collected as the GlcNAc fraction by centrifugation (10,000 × g, 10 min). The amount of GlcNAc was determined using a colorimetric method. Then, 500 µL of the GlcNAc fraction was mixed with 100 µL of 0.8 M borate buffer (pH 9.1, adjusted with KOH) and heated in boiling water for 3 min. After cooling in tap water, 3 mL of a coloring solution composed of 10 mg/mL p-dimethylaminobenzaldehyde and 125 mM hydrochloric acid in glacial acetic acid was added to the mixture and further incubated at 37°C for 20 min, giving rise to purple coloration. The absorbance of the reaction mixtures at 585 nm was measured, and the amount of GlcNAc in cultures (in micrograms of GlcNAc per gram culture) was estimated based on a standard curve generated using 0.05, 0.1, 0.15, and 0.2 µmol GlcNAc (Wako Pure Chemical, Osaka, Japan).

Determination of decarboxylase activity for arginine and ornithine

Agmatine and putrescine amounts were quantified using HPLC, as previously described (18). To monitor ADC and ODC activities, the products of enzymatic reactions were analyzed by HPLC. The reaction mixtures (500 µL) contained 0.1 mM substrate (arginine or ornithine) in 25 mM citrate buffer (pH 3.0, 4.0, 5.0, or 6.0). Each purified protein (5 µg) or crude cell extract was added to the mixture and then incubated at 30°C for 60 min. When PLP dependency was examined, 0.1 mM PLP was added to the reaction mixture. The reaction was terminated by the addition of 50 µL of 10% (wt/vol) TCA, and the reaction mixture was filtered through a 0.45 µm pore-size filter (Millex LH filter; Millipore, Bedford, MA, USA). The agmatine and putrescine contents in the filtrate were determined by HPLC to assess the ADC and ODC activity, respectively.

Determination of decarboxylase activity for phosphatidylserine

PSD activity was measured by detecting phosphatidylethanolamine. The reaction mixtures (250 µL) containing 0.2 mM phosphatidylserine, 0.2% Triton X, 25 mM phosphate buffer (pH 7.0), and 50 µL of enzyme sample were incubated at 30°C for 10 min. When PLP dependency was examined, 0.1 mM PLP was added to the reaction mixture. The reaction was terminated by adding 25 µL of 2 N KOH, following centrifugation (20,000 × g, 15 min, 4°C). The supernatant was obtained, and produced phosphatidylethanolamine was detected according to the phosphatidylethanolamine assay kit (Abcam, Cambridge, UK).

Natural ADC purification from A. oryzae for LC-MS/MS analysis

The rice koji obtained from 300 g of steamed rice was frozen with liquid nitrogen and resuspended in 200 mL of 50 mM phosphate buffer (pH 7.0), followed by crushing with mixer and sonication on ice. The disrupted hyphal sample was centrifuged (7,000 × g, 15 min), and the supernatant was obtained as crude cell extract. After the supernatant was obtained by centrifugation (7,000 × g, 15 min), ammonium sulfate was added to 60% saturation. The supernatant was obtained by centrifugation (15,000 × g, 20 min), and ammonium sulfate was added to 80% saturation. The precipitate was collected by centrifugation (15,000 × g, 20 min), dissolved in 50 mM phosphate buffer (pH 7.0), and dialyzed against the same buffer. The solution was applied to anion-exchange column, 5 mL Hitrap Q (Cytiva, Tokyo, Japan), followed by elution with a linear gradient of NaCl (0 to 1.0 M). The fractions with containing ADC activity were collected, dialyzed against 50 mM phosphate buffer (pH 7.0). The fractions were collected, concentrated with an Amicon Ultra-3K device (Millipore), applied to a Superdex 200 HR 10/30 gel filtration column (Cytiva), and eluted with the same buffer. Fractions with enzymatic activity were collected and concentrated. The concentrations of purified proteins were determined using the Bradford dye-binding assay (43) with bovine serum albumin (BSA) as a standard. The active fractions were concentrated and applied to an SDS-PAGE preparatory gel. The thick bands were cut out, dehydrated in acetonitrile, and alkylated by incubation with 55 mM iodoacetamide for 90 min. The product was digested with trypsin (Trypsin Gold, Promega, WI) overnight at 37°C, and desalted by Zip-Tip (Millipore). The isolated peptides were analyzed by liquid chromatography mass spectrometry (LC–MS) (Table S1) using an Orbitrap MS (Thermo Fisher Scientific K.K.) equipped with Advance UHPLC HTSxt-PAL (AMR), and Mascot was used as the sequence database search engine. The flow rate of LC was set to 300 nL/min with a gradient of 5% solution A (99.9% water and 0.1% formic acid) and 95% solution B (100% acetonitrile). Solution B was increased up to 35% at 40 min and further to 90% at 41 min. After holding for 5 min, the mixture was washed with 100% solution B was conducted for 60 min. The effluent from the HPLC was directly electrosprayed into the mass spectrometer. Peptides separated by LC were ionized at 1.7 kV in the positive ion mode. MS1 survey scans were acquired from 350 to 1800 m/z at a resolution of 60,000. Peptides were identified with a precursor mass tolerance of 10 ppm, and a fragment m/z tolerance of 0.8. MS/MS fragmentation spectra was searched with Proteome Discoverer 2.4 (Thermo Fisher Scientific K.K.). The results were analyzed by the Mascot™ v. 2.6 (Matrix Science, SC) using the following databases: MASCOT MS/MS ion search/via PD2.4, Uniprot AspergillusOryzae_RIB40_UP000006564, and Genome_NCBI.

Expression and purification of recombinant ADC of A. oryzae RIB40

The full-length Ao-adc1 gene was amplified by PCR from A. oryzae RIB genomic DNA using the specific primer pairs AO-CAA-1-Rv and AO-GAC-1-Fw (Table S4). The amplified DNA was digested with restriction enzymes, EcoRI and NdeI, and introduced into EcoRI and NdeI sites of plasmid pET28a (Takara Bio) (Table S3). The plasmid construction was confirmed by nucleotide sequence determination by using synthetic primers (1Ao-GCC-Fw2, 2Ao-GAC-Fw2, 3Ao-GAA-Fw2, 4Ao-CTC-Fw2, and 5Ao-GTG-Fw2 listed in Table S4). The resultant plasmid was designated as pET28a-Ao-ADC1. The plasmid was introduced into E. coli BL21 CodonPlus (DE3)-RIL cells, and the transformant cells were grown in the LB medium containing 20 µg/mL kanamycin and 20 µg/mL chloramphenicol at 37℃. The expression of the ADC with N-terminal His tags was induced by the addition of 1 mM IPTG. After further incubation at 37℃ or 20℃ for 4 hours, the cells were harvested by centrifugation, resuspended in 50 mM phosphate buffer [pH 7.0], and disrupted by sonication. After removing the cell debris by centrifugation, each supernatant was applied to a column packed with 5 mL of His60 Ni Superflow Resin (Takara Bio) and eluted stepwise with the 50 mM phosphate buffer [pH 7.0] supplemented with 100, 200, or 500 mM imidazole. Eluted protein was dialyzed against 50 mM phosphate buffer [pH 7.0]. The concentrations of purified proteins were determined using the Bradford dye-binding assay with BSA as a standard. To monitor purification, normal Tris-glycine SDS-PAGE was conducted and visualized by Coomassie brilliant blue staining. For small peptide monitoring Tris-tricine SDS-PAGE was performed according to the reported procedure (44) and visualized by silver staining.

Western blotting analysis

The effect of pH on ADC1 expression in A. oryzae RIB40 was examined by immunodetection with specific polyclonal antisera. The polyclonal antisera against Ao-ADC1 were obtained from Female New Zealand White rabbit (Japan SLC) and designed as anti-Ao-ADC1. A. oryzae RIB40 was cultured for 60 hours in solid state condition at the range of pH 3.0–6.0 by using 50 mM citric acid buffer. One gram of rice koji was frozen with liquid nitrogen and resuspended in 10 mL of 50 mM phosphate buffer (pH 7.0), followed by crushing with mixer and sonication on ice. Disrupted cells were centrifuged, and supernatants were obtained. These samples were subjected to electrophoresis in gels containing 0.1% SDS sulfate and 10% PAGE and then electroblotted onto polyvinylidene difluoride membranes (Bio-Rad, Tokyo, Japan). The membranes were blocked using 5% skimmed milk and incubated with rabbit anti-Ao-ADC1 serum at a 1:2,000 dilution subjected to SDS-PAGE, and the gel was stained with coomassie brilliant blue (CBB) to display the total protein. Western blotting was performed using antibodies that bind specifically. Detection of immunocomplexes was performed using Alexa Fluor 700 goat anti-rabbit IgG (Invitrogen, Carlsberg, CA), and the signal was visualized with an Odyssey infrared imaging system (LI-COR Biosciences, Lincoln, NE).

Phylogenetic analysis of amino acid sequences of phosphoserine dehydrogenase homologs

The phylogenetic analysis was conducted using the RAxML method provided by GenomeNet operated by the Kyoto University Bioinformatics Center (https://www.genome.jp/). Bootstrap resampling was performed 1,000 times, and only the values observed in more than 50% of the replicas are shown at the branching points. Scale bar indicates 1 substitution per amino acid.

ACKNOWLEDGMENTS

The authors want to acknowledge the scientific input and the continuous scientific support from the colleagues at the Department of Sciences, Kwansei-Gakuin University

SUPPLEMENTAL MATERIAL

Fig. S1 - aem.00294-24-s0001.eps
Mass spectra obtained from MSI analysis of steamed rice, rice koji prepared at pH 3.0, and rice koji prepared at pH 6.0.
Table S1 - aem.00294-24-s0002.pdf
Conditions used for LC/MS analysis.
Table S2 - aem.00294-24-s0003.pdf
Identified proteins from purified ADC fraction of A. oryzae RIB40.
Table S3 - aem.00294-24-s0004.pdf
Strains and plasmids used in this study.
Table S4 - aem.00294-24-s0005.pdf
Primers used in this study.
ASM does not own the copyrights to Supplemental Material that may be linked to, or accessed through, an article. The authors have granted ASM a non-exclusive, world-wide license to publish the Supplemental Material files. Please contact the corresponding author directly for reuse.

REFERENCES

1.
Minois N. 2014. “Molecular basis of the 'anti-aging' effect of Spermidine and other natural Polyamines - a mini-review”. Gerontology 60:319–326.
2.
Minois N, Rockenfeller P, Smith TK, Carmona-Gutierrez D. 2014. Spermidine feeding decreases age-related locomotor activity loss and induces changes in lipid composition. PLoS One 9:e102435.
3.
Gupta VK, Scheunemann L, Eisenberg T, Mertel S, Bhukel A, Koemans TS, Kramer JM, Liu KSY, Schroeder S, Stunnenberg HG, Sinner F, Magnes C, Pieber TR, Dipt S, Fiala A, Schenck A, Schwaerzel M, Madeo F, Sigrist SJ. 2013. Restoring polyamines protects from age-induced memory impairment in an autophagy-dependent manner. Nat Neurosci 16:1453–1460.
4.
Eisenberg T, Knauer H, Schauer A, Büttner S, Ruckenstuhl C, Carmona-Gutierrez D, Ring J, Schroeder S, Magnes C, Antonacci L, et al. 2009. Induction of autophagy by spermidine promotes longevity. Nat Cell Biol 11:1305–1314.
5.
Soda K, Dobashi Y, Kano Y, Tsujinaka S, Konishi F. 2009. Polyamine-rich food decreases age-associated pathology and mortality in aged mice. Exp Gerontol 44:727–732.
6.
Matsumoto M, Kurihara S, Kibe R, Ashida H, Benno Y. 2011. Longevity in mice is promoted by probiotic-induced suppression of colonic senescence dependent on upregulation of gut bacterial polyamine production. PLoS One 6:e23652.
7.
Piletz JE, Klenotich S, Lee KS, Zhu QL, Valente E, Collins MA, Jones V, Lee SN, Yangzheng F. 2013. Putative agmatinase inhibitor for hypoxic-ischemic new born brain damage. Neurotox Res 24:176–190.
8.
Laube G, Bernstein HG. 2017. Agmatine: multifunctional arginine metabolite and magic bullet in clinical neuroscience Biochem J 474:2619–2640.
9.
Akasaka N, Fujiwara S. 2020. The therapeutic and nutraceutical potential of agmatine, and its enhanced production using Aspergillus oryzae. Amino Acids 52:181–197.
10.
Battaglia V, Grancara S, Satriano J, Saccoccio S, Agostinelli E, Toninello A. 2010. Agmatine prevents the ca2+-dependent induction of permeability transition in rat brain mitochondria. Amino Acids 38:431–437.
11.
Law A, O’Donnell J, Gauthier S, Quirion R. 2002. Neuronal and inducible nitric oxide synthase expressions and activities in the hippocampi and cortices of young adult, aged cognitively unimpaired, and impaired long-evans rats. Neuroscience 112:267–275.
12.
Rushaidhi M, Collie ND, Zhang H, Liu P. 2012. Agmatine selectively improves behavioural function in aged male sprague-dawley rats. Neuroscience 218:206–215.
13.
Satriano J, Matsufuji S, Murakami Y, Lortie MJ, Schwartz D, Kelly CJ, Hayashi S, Blantz RC. 1998. Agmatine suppresses proliferation by frameshift induction of antizyme and attenuation of cellular polyamine levels. J Biol Chem 273:15313–15316.
14.
Nissim I, Horyn O, Daikhin Y, Chen P, Li C, Wehrli SL, Nissim I, Yudkoff M. 2014. The molecular and metabolic influence of long term agmatine consumption. J Biol Chem 289:9710–9729.
15.
Okamoto A, Sugi E, Koizumi Y, Yanagida F, Udaka S. 1997. Polyamine content of ordinary foodstuffs and various fermented foods. Biosci Biotechnol Biochem 61:1582–1584.
16.
Galgano F, Caruso M, Condelli N, Favati F. 2012. Focused review: agmatine in fermented foods. Front Microbiol 3:199.
17.
Bokulich NA, Ohta M, Lee M, Mills DA. 2014. Indigenous bacteria and fungi drive traditional Kimoto sake fermentations. Appl Environ Microbiol 80:5522–5529.
18.
Akasaka N, Kato S, Kato S, Hidese R, Wagu Y, Sakoda H, Fujiwara S. 2018. Agmatine production by Aspergillus oryzae is elevated by low pH during solid-state cultivation. Appl Environ Microbiol 84:e00722-18.
19.
Machida M, Asai K, Sano M, Tanaka T, Kumagai T, Terai G, Kusumoto K-I, Arima T, Akita O, Kashiwagi Y, et al. 2005. Genome sequencing and analysis of Aspergillus oryzae. Nature 438:1157–1161.
20.
Akao T, Yashiro I, Hosoyama A, Kitagaki H, Horikawa H, Watanabe D, Akada R, Ando Y, Harashima S, Inoue T, et al. 2011. Whole-genome sequencing of sake yeast Saccharomyces cerevisiae kyokai No.7. DNA Res 18:423–434.
21.
Valdés-Santiago L, Ruiz-Herrera J. 2013. Stress and polyamine metabolism in fungi. Front Chem 1:42.
22.
Kumar S, Saragadam T, Punekar NS. 2015. Novel route for agmatine catabolism in Aspergillus niger involves 4-guanidinobutyrase. Appl Environ Microbiol 81:5593–5603.
23.
Zhu MY, Iyo A, Piletz JE, Regunathan S. 2004. Expression of human arginine decarboxylase, the biosynthetic enzyme for agmatine. Biochim Biophys Acta 1670:156–164.
24.
Putri Wisman A, Tamada Y, Hirohata S, Fukusaki E, Shimma S. 2020. Metabolic visualization reveals the distinct distribution of sugars and amino acids in rice. Mass Spectrom (Tokyo) 9:A0089.
25.
Raetz CR, Kantor GD, Nishijima M, Newman KF. 1979. Cardiolipin accumulation in the inner and outer membranes of Escherichia coli mutants defective in phosphatidylserine synthetase. J Bacteriol 139:544–551.
26.
Dowhan W, Wickner WT, Kennedy EP. 1974. Purification and properties of phosphatidylserine decarboxylase from Escherichia coli. J Biol Chem 249:3079–3084.
27.
Moser R, Aktas M, Fritz C, Narberhaus F. 2014. Discovery of a bifunctional cardiolipin/phosphatidylethanolamine synthase in bacteria. Mol Microbiol 92:959–972.
28.
Calzada E, Avery E, Sam PN, Modak A, Wang C, McCaffery JM, Han X, Alder NN, Claypool SM. 2019. Phosphatidylethanolamine made in the inner mitochondrial membrane is essential for yeast cytochrome BC1 complex function. Nat Commun 10:1432.
29.
Zborowski J, Dygas A, Wojtczak L. 1983. Phosphatidylserine decarboxylase is located on the external side of the inner mitochondrial membrane. FEBS Lett 157:179–182.
30.
John RA. 1995. Pyridoxal phosphate-dependent enzymes. Biochim Biophys Acta 1248:81–96.
31.
Eliot AC, Kirsch JF. 2004. Pyridoxal phosphate enzymes: mechanistic, structural, and evolutionary considerations. Annu Rev Biochem 73:383–415.
32.
van Poelje PD, Snell EE. 1990. Pyruvoyl-dependent enzymes. Annu Rev Biochem 59:29–59.
33.
Watanabe Y, Watanabe Y, Watanabe S. 2020. Structural basis for phosphatidylethanolamine biosynthesis by bacterial phosphatidylserine decarboxylase. Structure 28:799–809.
34.
Choi JY, Duraisingh MT, Marti M, Ben Mamoun C, Voelker DR. 2015. From protease to decarboxylase: the molecular metamorphosis of phosphatidylserine decarboxylase. J Biol Chem 290:10972–10980.
35.
Ogunbona OB, Onguka O, Calzada E, Claypool SM. 2017. Multitiered and cooperative surveillance of mitochondrial Phosphatidylserine decarboxylase. Mol Cell Biol 37:e00049-17.
36.
Li QX, Dowhan W. 1988. Structural characterization of Escherichia coli phosphatidylserine decarboxylase. J Biol Chem 263:11516–11522.
37.
Kitano H, Kataoka K, Furukawa K, Hara S. 2002. Specific expression and temperature-dependent expression of the acid protease-encoding gene (pepA) in Aspergillus oryzae in solid-state culture (rice-Koji). J Biosci Bioeng 93:563–567.
38.
Wu WH, Morris DR. 1973. Biosynthetic arginine decarboxylase from Escherichia coli. J Biol Chem 248:1687–1695.
39.
Andréll J, Hicks MG, Palmer T, Carpenter EP, Iwata S, Maher MJ. 2009. Crystal structure of the acid-induced arginine decarboxylase from Escherichia coli: reversible decamer assembly controls enzyme activity. Biochemistry 48:3915–3927.
40.
Stim KP, Bennett GN. 1993. Nucleotide sequence of the adi gene, which encodes the biodegradative acid-induced arginine decarboxylase of Escherichia coli. J Bacteriol 175:1221–1234.
41.
Li QX, Dowhan W. 1988. Structural characterization of Escherichia coli phosphatidylserine decarboxylase. J Biol Chem 263:11516–11522.
42.
Wisman AP, Tamada Y, Hirohata S, Gomi K, Fukusaki E, Shimma S. 2020. Mapping haze-komi on rice koji grains using β-glucuronidase expressing Aspergillus oryzae and mass spectrometry imaging. J Biosci Bioeng 129:296–301.
43.
Bradford MM. 1976. A rapid and sensitive method for the Quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal Biochem 72:248–254.
44.
Schägger H, von Jagow G. 1987. Tricine-sodium dodecyl sulfate-polyacrylamide GEL electrophoresis for the separation of proteins in the range from 1 to 100 kDa. Anal Biochem 166:368–379.

Information & Contributors

Information

Published In

cover image Applied and Environmental Microbiology
Applied and Environmental Microbiology
Volume 90Number 521 May 2024
eLocator: e00294-24
Editor: Irina S. Druzhinina, Druzhinina, Royal Botanic Gardens, Surrey, United Kingdom
PubMed: 38624200

History

Received: 17 February 2024
Accepted: 23 March 2024
Published online: 16 April 2024

Permissions

Request permissions for this article.

Keywords

  1. Aspergillus oryzae
  2. arginine decarboxylase
  3. polyamine
  4. agmatine
  5. acid resistance mechanism

Contributors

Authors

Yui Murakami
Department of Biosciences, Graduate School of Science and Technology, Kwansei-Gakuin University, Gakuen-Uegahara, Sanda, Hyogo, Japan
Soichiro Ikuta
Department of Biosciences, School of Biological and Environmental Sciences, Kwansei-Gakuin University, Gakuen-Uegahara, Sanda, Hyogo, Japan
Wakao Fukuda
Department of Biosciences, School of Biological and Environmental Sciences, Kwansei-Gakuin University, Gakuen-Uegahara, Sanda, Hyogo, Japan
Naoki Akasaka
Department of Biosciences, Graduate School of Science and Technology, Kwansei-Gakuin University, Gakuen-Uegahara, Sanda, Hyogo, Japan
Laboratory for Circular Bioeconomy Development, Office of Society-Academia Collaboration for Innovation, Kyoto University, Kitashirakawa-Oiwakecho, Sakyo-ku, Kyoto, Japan
Jun-ichi Maruyama
Department of Biotechnology, The University of Tokyo, Bunkyo-ku, Tokyo, Japan
Collaborative Research Institute for Innovative Microbiology, The University of Tokyo, Bunkyo-ku, Tokyo, Japan
Shuichi Shinma
Department of Biotechnology, Graduate School of Engineering, Osaka University, Suita, Osaka, Japan
Osaka University Shimadzu Analytical Innovation Laboratory, Osaka University, Suita, Osaka, Japan
Institute for Open and Transdisciplinary Research Initiatives (OTRI), Osaka University, Suita, Osaka, Japan
Eiichiro Fukusaki
Department of Biotechnology, Graduate School of Engineering, Osaka University, Suita, Osaka, Japan
Osaka University Shimadzu Analytical Innovation Laboratory, Osaka University, Suita, Osaka, Japan
Institute for Open and Transdisciplinary Research Initiatives (OTRI), Osaka University, Suita, Osaka, Japan
Department of Biosciences, Graduate School of Science and Technology, Kwansei-Gakuin University, Gakuen-Uegahara, Sanda, Hyogo, Japan
Department of Biosciences, School of Biological and Environmental Sciences, Kwansei-Gakuin University, Gakuen-Uegahara, Sanda, Hyogo, Japan

Editor

Irina S. Druzhinina
Editor
Druzhinina, Royal Botanic Gardens, Surrey, United Kingdom

Notes

The authors declare no conflict of interest.

Metrics & Citations

Metrics

Note:

  • For recently published articles, the TOTAL download count will appear as zero until a new month starts.
  • There is a 3- to 4-day delay in article usage, so article usage will not appear immediately after publication.
  • Citation counts come from the Crossref Cited by service.

Citations

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. For an editable text file, please select Medlars format which will download as a .txt file. Simply select your manager software from the list below and click Download.

View Options

Figures and Media

Figures

Media

Tables

Share

Share

Share the article link

Share with email

Email a colleague

Share on social media

American Society for Microbiology ("ASM") is committed to maintaining your confidence and trust with respect to the information we collect from you on websites owned and operated by ASM ("ASM Web Sites") and other sources. This Privacy Policy sets forth the information we collect about you, how we use this information and the choices you have about how we use such information.
FIND OUT MORE about the privacy policy