Physarum polycephalum expresses two closely related, calcium-independent NOSs (nitric oxide synthases). In our previous work, we showed that both NOSs are induced during starvation and apparently play a functional role in sporulation. In the present study, we characterized the genomic structures of both Physarum NOSs, expressed both enzymes recombinantly in bacteria and characterized their biochemical properties. Whereas the overall genomic organization of Physarum NOS genes is comparable with various animal NOSs, none of the exon–intron boundaries are conserved. Recombinant expression of clones with various N-termini identified N-terminal amino acids essential for enzyme activity, but not required for haem binding or dimerization, and suggests the usage of non-AUG start codons for Physarum NOSs. Biochemical characterization of the two Physarum isoenzymes revealed different affinities for L-arginine, FMN and 6R-5,6,7,8-tetrahydro-L-biopterin.

Nitric oxide, a signalling and cytotoxic/cytoprotective agent, plays a key role in regulating the nervous, immune and cardiovascular systems of animals. It is synthesized by NOS (nitric oxide synthase; EC 1.14.13.39), which occurs as three isoforms and genes, i.e. NOS1, NOS2 and NOS3, also termed nNOS (neuronal NOS), iNOS (inducible NOS) and eNOS (endothelial NOS) respectively. In addition, a number of tissue- and development-specific splice variants encoding variant NOS proteins have been identified [14]. All three NOS isoenzymes are part of a complex network regulating nitric oxide production in health and disease (for recent reviews, see [59]).

NOSs oxidize L-arginine to L-citrulline and nitric oxide in an NADPH-dependent reaction via forming Nω-hydroxy-L-arginine as an enzyme-bound intermediate [10,11]. The enzymes consist of an N-terminal oxygenase domain binding haem and a C-terminal reductase domain binding FMN, FAD and NADPH. The two domains are linked through a calmodulin-binding motif. For activity, NOSs are required to form a homodimer leading to the formation of high-affinity binding sites for H4-bip (6R-5,6,7,8-tetrahydro-L-biopterin) and L-arginine in each oxygenase domain, thus allowing for the electron transfer between flavin and haem groups (reviewed in [12]). Detailed biochemical and crystallographic studies revealed structural features of the three isoenzymes underlying the observed differences in dimer formation and stability caused by L-arginine and H4-bip [13]. In addition to stabilizing the NOS dimer and affecting the haem spin state as well as the affinity of NOS for its substrate, H4-bip also serves as an electron/proton donor for nitric oxide formation, thereby giving rise to an H4-bip radical [1416]. Thus the role of H4-bip in NOS is unique as compared with aromatic amino acid hydroxylases, the classical H4-bip-dependent enzymes, where the cofactor activates oxygen by transferring two electrons, thereby being oxidized and enzymatically recycled after each catalytic turnover (reviewed in [14,15]). Under conditions of limited H4-bip availability, NADPH oxidation becomes uncoupled from nitric oxide synthesis, thus leading to O2/H2O2 production [11], a process involved in certain clinical settings such as endothelial dysfunction [15].

So far, the first and only NOS isolated and cloned from a non-animal species was found in Physarum polycephalum [17,18]. While many bacteria contain proteins with sequences similar to mammalian NOS oxygenase domains, these are not full–length NOSs but lack the reductase domain essential for nitric oxide production (for a review, see [11]). A defined functional role for these bacterial NOS oxygenase-like proteins, which are often termed bacterial NOS, is the nitration of tryptophan [19]. P. polycephalum is a unicellular model organism expressing two NOS isoforms (A and B) that share 82% amino acid identity and are expressed in parallel in the organism throughout various developmental stages. Starvation strongly induces expression of both NOSs, which appears to be a prerequisite to obtain sporulation competence, a process that also involves nitric oxide-sensitive guanylate cyclase [18]. The overall sequence identity to mammalian NOSs is less than 39%. Nevertheless, binding motifs for flavins, NADPH, H4-bip, calmodulin and caveolin are highly conserved in Physarum NOSs and both proteins resemble mammalian iNOS lacking the spacer sequence conferring calcium dependence [18].

In the present study, we characterized the genomic structures of both Physarum NOSs and set up bacterial overexpression of recombinant proteins. This allowed purification and biochemical characterization of Physarum NOSs as well as identification of the minimal N-terminal sequence being essential for activity.

Cultivation of P. polycephalum

Strain M3b, a Wis 1 isolate used initially for purifying and cloning of Physarum NOS, was grown in a semi-defined medium as detailed previously [18]. The apogamic haploid strain LU352 was kindly provided by Professor Dr Wolfgang Marwan (Max Planck Institute for Dynamics of Complex Technical Systems, Magdeburg, Germany) and grown as described in [20]. Amoebae were generated and grown as described in [21].

RNA isolation and quantitative PCR

RNA from LU352 amoebae and starved macroplasmodia was isolated using the RNeasy plant mini kit (Qiagen). Physarum NOS form A and B mRNAs were quantified using Taqman technology. Sequences for probes and primers were for NOS form A: 5′-ACGCGCACACAGCCAAGAAACG-3′ (probe), 5′-CATCCCCGAAACTGTTGCTC-3′ (forward primer), 5′-GCAGTCCGTGGTAGCAACCT-3′ (reverse primer); for NOS form B: 5′-CGCAGCGTGCTATCAGCCCAGA-3′ (probe), 5′-CCAGAAGTACACTATTCCATCAGAAATC-3′ (forward primer), 5′-GCGGCAATCAGATGTGGATA-3′ (reverse primer). For a reference, 19 S RNA was quantified as described in [18].

DNA isolation

DNA from frozen M3b microplasmodia suspended in lysis buffer was isolated using the DNeasy plant maxi kit (Qiagen) according to the manufacturer's instructions and precipitated using sodium acetate and ethanol. DNA was further purified using the UltraPure reagent (Gibco) and precipitated once more with sodium acetate and ethanol.

PCR, genome walking, cloning and library screening

Genomic structures of Physarum NOSs were elucidated using a combination of various techniques. First, PCR primer pairs close to exon/exon boundaries of mammalian NOSs were synthesized using Physarum NOS cDNA sequences [18] [GenBank® Nucleotide Sequence Database accession numbers AF145041 (form A) and AF145040 (form B)]. Those regions that could not be identified by this approach were identified by genome walking using the Universal GenomeWalker kit from Clontech according to the manufacturer's instructions. Herculase Hotstart DNA polymerase (Stratagene) generating an A-overhang was used for PCR and genome walking, and the generated products were then cloned into TOPO TA cloning vector (Invitrogen). Finally, most gaps could be closed by screening a genomic library [produced by Professor Tim Burland (McArdle Laboratory for Cancer Research, University of Wisconsin, Madison, WI, U.S.A.) and kindly provided by Professor Jonatha Gott (Center for RNA Molecular Biology, Case Western Reserve University, Cleveland, OH, U.S.A.)] using standard procedures. The library had been prepared from the LU352 strain by the customer service of Stratagene by using the LambdaZap vector.

Sequencing and data analysis

Sequencing of clones and plasmids was done by the custom service of Microsynth (Balgach, Switzerland). Sequence data were analysed using the Wisconsin Sequence Analysis Package version 10.3 by the Genetics Computer Group (Accelrys, Cambridge, U.K.).

Plasmids for the overexpression of Physarum NOSs

Physarum calmodulin cDNA (GenBank® accession number AB022702) was amplified using PCR and cloned into site 2 of the pET-Duet1 vector (Novagen) using EcoRV/XhoI restriction sites. Physarum NOS form A cDNA (AF145041, cDNA library clone) was then cloned into site 1 of the pET-Duet1 vector using SalI/NotI restriction digestion. This sequence included 104 amino acids upstream of the putative start methionine [18], 11 amino acids stemming from the original Bluescript SK vector, and 25 amino acids from the pET-Duet1 vector including a His tag. Alternatively, Physarum NOS form B (AF145040) was cloned into the second cloning site of the pET-Duet1 vector using SacI/NotI restriction digestion. This cDNA started at the putative start methionine [18] and also contained an N-terminal His tag. Various N-termini of Physarum NOS form A were generated by the introduction of an additional SalI site at the desired position (QuikChange® II kit; Stratagene), cutting by SalI and re-ligating. Mutations were introduced by site-directed mutagenesis (QuikChange® II kit). For elongating the N-terminus of Physarum NOS form B, 155 bp of the genomic sequence 5′ to the putative start ATG were introduced using the Infusion cloning kit (Clontech). This region was confirmed to be identical with the cDNA 5′-region generated by RACE (rapid amplification of cDNA ends) PCR (see below). This clone was further modified by introducing an additional SacI restriction site at the desired position (QuikChange® II kit), cutting by SacI and re-ligating, thus leading to recombinant proteins with various N-termini. In addition, we introduced an N-terminal StrepTag [22] using the QuikChange® II protocol into one form of physnosa (A1) and its almost inactive mutant (A1LD) to allow the preparation of homogeneous, active proteins.

5′ Elongation of cDNA clones

In order to elongate 5′ sequences of the previously published Physarum NOS form A and B cDNA clones [18], the SMART RACE cDNA amplification kit (Clontech) was used.

Bacterial overexpression of Physarum NOSs and preparation of homogenates

For overexpression, pET-Duet1 expression plasmids (see above), grown in NovaBlue (Novagen) cells and purified by the SNAP midi prep plasmid preparation kit (Invitrogen), were transformed into TUNER DE3 cells (Novagen). Bacterial starter cultures (5 ml) from single colonies were grown in LB (Luria–Bertani) broth (Difco) with ampicillin (50 μg/ml) at 37 °C and 220 rev./min to an attenuance at 600 (D600) of 0.8, and then pelleted and diluted in 15 ml of fresh TB (Terrific broth) without ampicillin but supplemented with 1.5% (v/v) ethanol in order to increase recovery of soluble protein by induction of a heat-shock response [23]. Aliquots for determining NOS activity at zero time were taken and cultures were supplemented with 450 μM 5-aminolaevulinic acid (Sigma A-7793) and 3 μM riboflavin (Sigma R-4500), in order to provide sufficient haem and flavins, and 50 μM IPTG (isopropyl β-D-thiogalactoside). Bacteria were grown for a further 24 h at 20 °C at 220 rev./min in the dark for NOS expression. For purification, this protocol was scaled up to 30 ml starter cultures and 400 ml expression cultures. TB was used instead of LB for expression cultures. Bacteria from these cultures were resuspended in 40 ml of buffer A [50 mM Tris/HCl, pH 8.0, containing 10% (v/v) glycerol, 25 μM FAD, 25 μM FMN, 50 μM L-arginine, 5 μM H4-bip (Schircks Laboratories, Jona, Switzerland), 5 mM DTE (dithioerythritol), 1 mM PMSF and 0.2 mg/ml lysozyme]. After incubation for 20 min at 25 °C at 200 rev./min, cells were homogenized using a French press [20000 lbf/in2 (1 lbf/in2=6.9 kPa) twice] and mixed with 110 ml of buffer B [50 mM Tris/HCl, pH 8.0, containing 10% (v/v) glycerol, 100 μM L-arginine, 10 μM H4-bip and 5 mM DTE].

Purification of recombinant Physarum NOSs

All steps were carried out at 4 °C and all elution buffers contained 100 μM L-arginine and 10 μM H4-bip. Protein was precipitated using 35.6 g of ammonium sulfate (45% saturation) and stirring for 1 h followed by centrifugation at 48000 g for 20 min. The pellet was then resuspended in 20 ml of buffer B by stirring for 1 h and undissolved material was collected by centrifugation at 48000 g for 20 min. The supernatant was dialysed twice for 2 h against 1 litre of buffer B using a 50K membrane (Pierce). The dialysate was adjusted to 0.15 M NaCl and then incubated (at 4 °C) with 2′,5′-ADP-Sepharose 4B (Amersham Biosciences) for 1 h using a rocking mixer at a ratio of 1 part dialysate protein to 2 parts ADP-Sepharose (w/w). After washing twice with 10 ml of buffer B containing 0.15 M NaCl and washing once with 10 ml of buffer B, protein was eluted from ADP-Sepharose with 1 ml of buffer B containing 10 mM NADPH. Protein determination at various steps was performed using the Bradford Protein Assay reagent from Bio-Rad. To obtain homogeneous active protein, the dimeric fractions of NOS were then collected by gel filtration on a Superose 12 10/300 GL column (GE Healthcare), and were finally purified using the StrepTag affinity protocol [22].

Native Physarum NOS

Native Physarum NOS was purified from microplasmodia as described in [18].

NOS assay

Radiometric citrulline NOS assay was carried out as described previously [17] using buffer B (see above) supplemented with 2 mM NADPH as the assay buffer. Briefly, bacterial homogenates (bacteria from 2 ml of culture were diluted in 700 μl of buffer A, incubated for 10 min at 25 °C and 1000 rev./min in an Eppendorf thermomixer and finally mixed for 10 s with an Ultra-Turrax homogenizer) or partially purified NOS (5–20 μl per assay corresponding to 10–15 μg of protein in the case of bacterial homogenates or ∼1 μg of protein when using purified enzyme), were incubated in assay buffer containing 40000 c.p.m. of HPLC-purified 3H-L-arginine (Amersham Biosciences) in a final volume of 200 μl for 10 min at 25 °C on an Eppendorf thermomixer. The reaction was then stopped, separated on Dowex 50 W columns and the flow-through was subjected to scintillation counting. In some assays, concentrations of cofactors or substrate were varied. IC50, EC50 and KM values were calculated with SigmaPlot 9.0 software (Systat Software).

For comparison of nitric oxide formation with NADPH consumption, a kinetic microtitre-plate assay based on measurement of nitric oxide formation by binding to myoglobin (monitored by UV absorption at 405 nm) combined with measurement of NADPH consumption (monitored at 340 nm) was performed at 25 °C as described in [24], using a PowerWaveX thermostatically controlled kinetic microplate reader (BioTek Instruments, Winooski, VT, U.S.A.).

Gel electrophoresis

Proteins were separated on SDS/polyacrylamide gels. Gels were stained using Coomassie Blue or silver stain according to standard procedures and scanned using an ImageScanner (Amersham Biosciences). For quantification, gels were stained with Deep Purple Total Protein Stain (Amersham Biosciences) and fluorescence was scanned by a Typhoon 9410 scanner (Amersham Biosciences) at 532 nm and evaluated using ImageQuant software (Amersham Biosciences).

Determination of haem content by HPLC

Haem content of pure forms A1 and A1LD was checked by HPLC according to the method of Bonkovsky et al. [25] by reversedphase HPLC separation of porphyrins and UV detection at 405 nm. The suitability of the application of this method for the analysis of protein-associated haem was confirmed with equine myoglobin (Serva, Heidelberg, Germany) yielding 0.9±0.1 haem per myoglobin.

Expression of Physarum NOS form A and B mRNA

We first wanted to clarify whether the two highly related Physarum NOS isoforms [18] are the products of two genes or of two alleles of the same gene, as was shown for the Physarum hapP gene [26]. For this purpose, we used the haploid Physarum strain LU352 and checked for the expression of Physarum NOS form A and B mRNA. Both mRNAs were expressed in LU352 macroplasmodia starved for 5 days in a way comparable with the diploid M3b strain, which we used in our initial studies [17,18]. Moreover, LU352 amoebae also expressed both mRNAs (Figure 1).

Quantification of Physarum NOS form A and form B mRNA levels

Figure 1
Quantification of Physarum NOS form A and form B mRNA levels

Total RNA was isolated from starved macroplasmodia of the diploid M3b and the haploid LU352 strain as well as from LU352 amoebae, and the two NOS forms were quantified by Taqman technology in relation to 19 S RNA. The means±S.D. for three independent experiments are shown. The dashed line indicates the detection limit of the method. Bar 1, M3b macroplasmodia, Physarum NOS form A; bar 2, M3b macroplasmodia, Physarum NOS form B; bar 3, LU352 macroplasmodia, Physarum NOS form A; bar 4, LU352 macroplasmodia, Physarum NOS form B; bar 5, LU 352 amoebae, Physarum NOS form A; bar 6, LU352 amoebae, Physarum NOS form B.

Figure 1
Quantification of Physarum NOS form A and form B mRNA levels

Total RNA was isolated from starved macroplasmodia of the diploid M3b and the haploid LU352 strain as well as from LU352 amoebae, and the two NOS forms were quantified by Taqman technology in relation to 19 S RNA. The means±S.D. for three independent experiments are shown. The dashed line indicates the detection limit of the method. Bar 1, M3b macroplasmodia, Physarum NOS form A; bar 2, M3b macroplasmodia, Physarum NOS form B; bar 3, LU352 macroplasmodia, Physarum NOS form A; bar 4, LU352 macroplasmodia, Physarum NOS form B; bar 5, LU 352 amoebae, Physarum NOS form A; bar 6, LU352 amoebae, Physarum NOS form B.

Close modal

Genomic structures of both Physarum NOSs

We then studied the genomic structures of both Physarum NOSs. As can be seen from Figure 2(A), 25 exons could be identified for Physarum NOS form B. Form A has two additional 5′ exons as was determined by extending the cDNA into the 5′ direction, a strategy that was not successful for the form B cDNA where only a shorter 5′ extension could be achieved starting within an exon (Figure 2A). On the other hand, the exact number of exons for NOS form A could not be determined since some internal genomic regions were inaccessible to sequencing (Figure 2A). However, since all of the 21 exon/intron boundaries that could be determined were conserved between the two Physarum NOSs, it is conceivable that the lacking 5 boundaries are also conserved. We therefore concluded that the two genes consist of 27 exons. For comparison, the gene structures of iNOS from human, chicken and trout are similarly organized into 27, 28 and 27 exons respectively (Figure 2A). However, none of the exon/intron boundaries are conserved between Physarum and the animal iNOSs (Figure 2A). Moreover, the size of introns in the Physarum genes is considerably smaller compared with animal NOSs, particularly the human iNOS gene (Figure 2B). As in other animal iNOSs, the reading frame of the Physarum NOSs starts in exon 3 (Figure 2A). The current version of the two gene structures was submitted to GenBank® Nucleotide Sequence Database (DQ835529, DQ845107 and DQ835525 for NOS form A and DQ835527 for NOS form B). The extended cDNA sequences were also submitted: DQ835526 for NOS form A and DQ825528 for NOS form B. The final gene structures will become available soon through the ongoing Physarum genome project (http://www.genome.gov/12511858 and http://genome.wustl.edu/genome.cgi?GENOME=Physarum%20polycephalum).

Genomic organization of Physarum NOSs

Figure 2
Genomic organization of Physarum NOSs

(A) Exons in cDNAs of human iNOS (bar 1, 27 exons; GenBank® accession numbers for genomic sequence: NT_010799.14; cDNA: NM_000625), chicken iNOS (bar 2, 28 exons; genomic: NW_060634.1|Gga19_WGA443_1; cDNA: NM_204961), trout iNOS (bar 3, 27 exons; genomic: AJ295231; cDNA: AJ295230), Physarum NOS form A [bar 4, 27 exons deduced from the comparison of forms A and B; see the main text for details; genomic: DQ835529 (5′-region, 6 exons), DQ845107 (internal region, 5 exons), DQ835525 (3′-region, 9 exons); cDNA: DQ835526] and form B (bar 5, 27 exons deduced from the comparison of forms A and B; see the main text for details; genomic: DQ835527; cDNA: DQ835528). Lines below each bar indicate the translated region of the cDNA. Positions of various binding sites are marked as have been published previously [4347]. For Physarum NOS form A, some internal exon/intron boundaries could not be determined by sequencing but were deduced from form B. This region is shown in white and the five deduced exon/exon boundaries are indicated by a thin line (bar 4). Also, exon 18 could not be fully sequenced. For Physarum NOS form B, the 5′-region could only be partially extended and starts within an exon (bar 5, indicated by interrupting the margin at the 5′-end), which was assigned number 3 as deduced from homology to NOS form A. (B) Intron sizes (bp) [GenBank® accession numbers for genomic and cDNA sequences are given under (A)] from iNOS genes of human (huminos), chicken (galinos), trout (oncinos) and the two Physarum NOSs (physnosa; physnosb). ns, not sequenced; –, not identified.

Figure 2
Genomic organization of Physarum NOSs

(A) Exons in cDNAs of human iNOS (bar 1, 27 exons; GenBank® accession numbers for genomic sequence: NT_010799.14; cDNA: NM_000625), chicken iNOS (bar 2, 28 exons; genomic: NW_060634.1|Gga19_WGA443_1; cDNA: NM_204961), trout iNOS (bar 3, 27 exons; genomic: AJ295231; cDNA: AJ295230), Physarum NOS form A [bar 4, 27 exons deduced from the comparison of forms A and B; see the main text for details; genomic: DQ835529 (5′-region, 6 exons), DQ845107 (internal region, 5 exons), DQ835525 (3′-region, 9 exons); cDNA: DQ835526] and form B (bar 5, 27 exons deduced from the comparison of forms A and B; see the main text for details; genomic: DQ835527; cDNA: DQ835528). Lines below each bar indicate the translated region of the cDNA. Positions of various binding sites are marked as have been published previously [4347]. For Physarum NOS form A, some internal exon/intron boundaries could not be determined by sequencing but were deduced from form B. This region is shown in white and the five deduced exon/exon boundaries are indicated by a thin line (bar 4). Also, exon 18 could not be fully sequenced. For Physarum NOS form B, the 5′-region could only be partially extended and starts within an exon (bar 5, indicated by interrupting the margin at the 5′-end), which was assigned number 3 as deduced from homology to NOS form A. (B) Intron sizes (bp) [GenBank® accession numbers for genomic and cDNA sequences are given under (A)] from iNOS genes of human (huminos), chicken (galinos), trout (oncinos) and the two Physarum NOSs (physnosa; physnosb). ns, not sequenced; –, not identified.

Close modal

Recombinant overexpression of Physarum NOSs in Escherichia coli

Our next goal was to set up efficient recombinant overexpression of Physarum NOSs in E. coli. In our previous work, we could functionally express Physarum NOS form A in a baculovirus system [18], but expression levels were low and did not allow purification of recombinant protein. It is known that high-level expression of mouse iNOS requires co-expression of calmodulin [27]. We therefore cloned Physarum calmodulin and co-expressed it together with Physarum NOS forms A and B respectively using the pET-Duet1 vector. In contrast with the original Physarum NOS form A cDNA library clone (AF145041) that still contained 104 additional amino acids 5′ to the putative start methionine [18] and a variant A1, which contained only 29 amino acids 5′ to the putative start, the clone starting at the predicted methionine (A2) had no enzyme activity (Figure 3A), although it expressed the same amount of protein (Figure 3D). Elimination of vector-derived sequences including the His tag resulted in disappearance of both protein and activity, presumably due to proteolytic degradation (results not shown). Systematically shortening the N-terminus 5′ to the assumed start methionine then revealed that 13 amino acids were sufficient to achieve a clone with enzyme activity (variant A8), which dropped dramatically when one amino acid (leucine) was deleted (variant A9) and became undetectable when deleting a further amino acid (variant A3) (Figure 3A). Exchanging this leucine residue in A1 for aspartic acid or for glycine (variants A1LD and A1LG) destroyed enzyme activity almost completely (Figure 3A), whereas it did not affect the amount of overexpressed protein (Figure 3B), and did not affect either the haem content of homogeneously purified proteins (see below) or the monomer/dimer ratio on gel filtration (results not shown). Expressing Physarum NOS form B using the predicted start methionine (clone B) also resulted in an inactive protein, whereas addition of 50 amino acids 5′ to this methionine (clone B+) yielded active protein. Varying this N-terminus showed that only four additional amino acids were sufficient for full activity of Physarum NOS form B (Figure 3A). As in form A, removal of a hydrophobic amino acid (leucine for truncation from B6 to B7 and isoleucine for truncation from B9 to B8) leads to a decrease in activity, although the amount of NOS protein in inactive clones is comparable with that of active clones (Figure 3E). These results clearly indicated that the N-terminus of both Physarum NOSs as predicted previously [18] may be too short. In line with this, comparison of the proteins expressed from the A and B clones and their variants with native Physarum NOS showed that the molecular masses of several of these proteins were smaller than the native protein, which is a mixture of both isoforms [18] (Figure 3C).

Overexpression of recombinant Physarum NOSs

Figure 3
Overexpression of recombinant Physarum NOSs

(A) N-termini of various clones of Physarum NOS form A and form B and their activity in whole homogenates. Values are means for five to ten independent experiments (±S.D). The previously predicted start of translation is underlined [18]. All overexpressed proteins start with a His tag sequence (MGSSHHHHHHSQDPNSSSARLQVD), indicated by an ellipsis. Clone A has an additional GIDKLDIEFRSKG sequence from the Bluescript SK vector inserted between the His tag and the Physarum NOS sequence. aa, amino acids. (B) Overexpression of A1 and its A1LD mutant. Homogenates (15 μg) harvested at times 0 and 24 h after induction were separated on SDS/6% polyacrylamide gels and stained with Coomassie Blue. For the induced cultures of A1 and A1LD, samples from two different expressions are shown. (C) Comparison of overexpressed Physarum NOS from clones A, A1 and B+ (3 μg of homogenate protein per lane) with native NOS purified from Physarum microplasmodia (0.3 μg per lane) [18]. Bands were detected by silver staining. Sequencing demonstrated that the second band in native Physarum NOS is a proteolytic fragment of the first one, and that both bands contained a mixture of Physarum NOS form A and Physarum NOS form B [18]. (D) Silver-stained gel of various A clones (homogenates, 3 μg of protein per lane). (E) Silver-stained gel of various B clones (homogenates, 3 μg per lane). Molecular masses (kDa) of several recombinant proteins were calculated for proteins containing the His tag and are given below the lanes of (CE).

Figure 3
Overexpression of recombinant Physarum NOSs

(A) N-termini of various clones of Physarum NOS form A and form B and their activity in whole homogenates. Values are means for five to ten independent experiments (±S.D). The previously predicted start of translation is underlined [18]. All overexpressed proteins start with a His tag sequence (MGSSHHHHHHSQDPNSSSARLQVD), indicated by an ellipsis. Clone A has an additional GIDKLDIEFRSKG sequence from the Bluescript SK vector inserted between the His tag and the Physarum NOS sequence. aa, amino acids. (B) Overexpression of A1 and its A1LD mutant. Homogenates (15 μg) harvested at times 0 and 24 h after induction were separated on SDS/6% polyacrylamide gels and stained with Coomassie Blue. For the induced cultures of A1 and A1LD, samples from two different expressions are shown. (C) Comparison of overexpressed Physarum NOS from clones A, A1 and B+ (3 μg of homogenate protein per lane) with native NOS purified from Physarum microplasmodia (0.3 μg per lane) [18]. Bands were detected by silver staining. Sequencing demonstrated that the second band in native Physarum NOS is a proteolytic fragment of the first one, and that both bands contained a mixture of Physarum NOS form A and Physarum NOS form B [18]. (D) Silver-stained gel of various A clones (homogenates, 3 μg of protein per lane). (E) Silver-stained gel of various B clones (homogenates, 3 μg per lane). Molecular masses (kDa) of several recombinant proteins were calculated for proteins containing the His tag and are given below the lanes of (CE).

Close modal

Checking the elongated Physarum NOS cDNAs showed that form A cDNA (DQ835526) contained a stop codon 516 bp upstream of the predicted start ATG but no additional methionine codon, whereas the form B cDNA (DQ835528) had no stop codon over the whole cloned 740 bp region 5′ to the putative start codon but encoded for an additional methionine 69 amino acids upstream of the predicted start [18] (Figure 4A). Starting with this methionine, the predicted form B protein (Figure 4B) is still smaller than the native Physarum NOS (Figure 3C). Searching the cDNA sequence upstream of the putative start codon showed that both Physarum NOSs have several putative non-AUG start codons [28,29]. An isoleucine residue conserved in both Physarum NOSs is found 88 (form A) or 90 (form B) amino acids upstream (Figure 4A). However, the Kozak sequence for these codons is less optimal [30] than that of other nearby residues (Figure 4B): for Physarum NOS form A cDNA, this is the isoleucine codon (ATT) at amino acid position –97; for Physarum NOS form B, an alternative start codon (ATT) with a nearly optimal context is encoding isoleucine in position –108. These sites yield proteins of nearly identical molecular mass (Figure 4B) that are comparable with that of native NOS (Figure 3C).

Putative translation start sites in Physarum NOSs

Figure 4
Putative translation start sites in Physarum NOSs

(A) Amino acid sequence predicted from 5′ extended cDNA clones of Physarum NOS form A (physnosa, GenBank® accession number DQ835526) and form B (physnosb, DQ835528). Conserved residues are shown in boldface. The previously predicted start of the protein [18] is framed. An additional methionine in form B is marked by a grey arrow, and putative alternative starts of the reading frames are indicated by black arrows. Numbers in italics refer to amino acid positions related to the previously predicted reading frames [18], and normal numbers refer to amino acids deduced from the 5′ extended cDNA clones (DQ835526 and DQ835528). The asterisk denotes the stop found in the Physarum NOS form A cDNA. (B) Molecular mass for Physarum NOSs with various starts and the Kozak sequences of their putative start codons. Putative start codons are written in lower-case letters, whereas the context for translation is given in upper-case letters. The nucleotide positions as well as the amino acid positions given in parentheses refer to GenBank® sequences DQ835526 and DQ835528. Amino acid positions not in brackets refer to the previously predicted reading frame [18] starting with methionine 1. The asterisk indicates previously published data [18].

Figure 4
Putative translation start sites in Physarum NOSs

(A) Amino acid sequence predicted from 5′ extended cDNA clones of Physarum NOS form A (physnosa, GenBank® accession number DQ835526) and form B (physnosb, DQ835528). Conserved residues are shown in boldface. The previously predicted start of the protein [18] is framed. An additional methionine in form B is marked by a grey arrow, and putative alternative starts of the reading frames are indicated by black arrows. Numbers in italics refer to amino acid positions related to the previously predicted reading frames [18], and normal numbers refer to amino acids deduced from the 5′ extended cDNA clones (DQ835526 and DQ835528). The asterisk denotes the stop found in the Physarum NOS form A cDNA. (B) Molecular mass for Physarum NOSs with various starts and the Kozak sequences of their putative start codons. Putative start codons are written in lower-case letters, whereas the context for translation is given in upper-case letters. The nucleotide positions as well as the amino acid positions given in parentheses refer to GenBank® sequences DQ835526 and DQ835528. Amino acid positions not in brackets refer to the previously predicted reading frame [18] starting with methionine 1. The asterisk indicates previously published data [18].

Close modal

Enzymatic properties of recombinant Physarum NOSs

We then characterized enzymatic properties of recombinant Physarum NOSs. For this purpose, we used proteins expressed from the A1 and B+ clone partially purified by ammonium sulfate precipitation and subsequent ADP-Sepharose-affinity chromatography. This procedure yielded NOS with approx. 80% purity (cf. Figure 5A, last lane). The specific activity was 244±47 nmol·mg−1·min−1 (form A, mean for three purifications±S.D.) and 253±47 nmol·mg−1·min−1 (form B, mean for four 4 purifications±S.D.). Figure 5 shows a typical purification of A1 recombinant protein. While both Physarum NOSs were comparably sensitive to NOS inhibitors and similarly dependent on NADPH and FAD, and had similar dimer/monomer ratios of approx. 1:1 on gel filtration (Figures 6A and 6B), the KM for L-arginine and the EC50 for FMN and H4-bip were clearly higher for NOS form A than for the form B protein (Table 1). A combined nitric oxide formation (determined by binding to myoglobin) and NADPH molecules consumption assay [24] showed partial decoupling of the enzyme preparations, with 5.1±1.6 NADPH molecules consumed in the formation of 1 nitric oxide molecule by Physarum NOS A and 5.9±1.5 NADPH molecules consumed per nitric oxide molecule in Physarum NOS B (mean±S.D. for three determinations).

Partial purification of recombinant Physarum NOS form A (clone A1)

Figure 5
Partial purification of recombinant Physarum NOS form A (clone A1)

(A) Coomassie-Blue-stained SDS/polyacrylamide gel (6–17% gradient) showing partial purification of recombinant Physarum NOS form A clone A1. The amount of total protein loaded per lane was 20 μg. Homogenate, whole bacterial homogenate; supernatant, supernatant after loading the 2′,5′-ADP-Sepharose 4B column; wash, first wash of 2′,5′-ADP-Sepharose 4B with buffer B; eluate, protein eluted from 2′,5′-ADP-Sepharose 4B with buffer B containing 10 mM NADPH. The ammonium sulfate precipitation step is not shown. (B) Activities, yield and purification factor of Physarum NOS form A clone A1 protein purified from 400 ml of bacterial culture (homogenate) by ammonium sulfate precipitation (AS precipitation) and 2′,5′-ADP-Sepharose 4B affinity chromatography (2 5-ADP).

Figure 5
Partial purification of recombinant Physarum NOS form A (clone A1)

(A) Coomassie-Blue-stained SDS/polyacrylamide gel (6–17% gradient) showing partial purification of recombinant Physarum NOS form A clone A1. The amount of total protein loaded per lane was 20 μg. Homogenate, whole bacterial homogenate; supernatant, supernatant after loading the 2′,5′-ADP-Sepharose 4B column; wash, first wash of 2′,5′-ADP-Sepharose 4B with buffer B; eluate, protein eluted from 2′,5′-ADP-Sepharose 4B with buffer B containing 10 mM NADPH. The ammonium sulfate precipitation step is not shown. (B) Activities, yield and purification factor of Physarum NOS form A clone A1 protein purified from 400 ml of bacterial culture (homogenate) by ammonium sulfate precipitation (AS precipitation) and 2′,5′-ADP-Sepharose 4B affinity chromatography (2 5-ADP).

Close modal

Gel filtration of purified Physarum NOSs

Figure 6
Gel filtration of purified Physarum NOSs

Physarum NOS isoforms (clones A1 and B+ respectively) carrying an N-terminal Strep tag were expressed together with Physarum calmodulin in E. coli using the pETDuet1 vector (see the Experimental section for details). A 200 μl portion of purified protein was then loaded on to a Superose 12 10/300 GL gel filtration column and eluted with buffer B containing 0.15 M NaCl. Fractions were collected and NOS activity was determined by the radiometric citrulline assay. The solid line shows UV absorption (OD) at 280 nm, the dashed line shows NOS activity. Arrows indicate the elution volume of protein standards for gel filtration (thyreoglobulin, 690 kDa; β-amylase, 200 kDa; albumin, 67 kDa). (A) Physarum NOS A; (B) Physarum NOS B; (C) silver-stained SDS gel showing the purity of Strep-tagged Physarum NOS A.

Figure 6
Gel filtration of purified Physarum NOSs

Physarum NOS isoforms (clones A1 and B+ respectively) carrying an N-terminal Strep tag were expressed together with Physarum calmodulin in E. coli using the pETDuet1 vector (see the Experimental section for details). A 200 μl portion of purified protein was then loaded on to a Superose 12 10/300 GL gel filtration column and eluted with buffer B containing 0.15 M NaCl. Fractions were collected and NOS activity was determined by the radiometric citrulline assay. The solid line shows UV absorption (OD) at 280 nm, the dashed line shows NOS activity. Arrows indicate the elution volume of protein standards for gel filtration (thyreoglobulin, 690 kDa; β-amylase, 200 kDa; albumin, 67 kDa). (A) Physarum NOS A; (B) Physarum NOS B; (C) silver-stained SDS gel showing the purity of Strep-tagged Physarum NOS A.

Close modal
Table 1
Enzymatic properties of recombinant Physarum NOSs

Partially purified recombinant Physarum NOS form A (clone A1) and form B (clone B+), each of them with an activity of approx. 250 nmol·mg−1·min−1 (see text for data), were used for measuring substrate, cofactor and inhibitor effects. L-NIL, L-N6(1-iminoethyl)-lysine; L-NNA, Nω-nitro-L-arginine; L-NMMA, Nω-methyl-L-arginine. Values are means±S.D. for three to five independent experiments. Values in boldface are significantly different (Student's t test) between NOS form A and form B: KML-arginine, P<0.001; EC50 H4-bip, P<0.001; EC50 FMN, P<0.02.

Concentration (μM)
Physarum NOS APhysarum NOS B
KML-arginine 66.7±14.5 26.8±9.7 
EC50 H4-bip 7.9 ±1.2 0.7±0.1 
EC50 NADPH 57.2±6.5 60.3±13.8 
EC50 FAD 3.9±1.6 2.1±1.6 
EC50 FMN 1.7±0.6 0.2±0.2 
IC50L-NIL 81.3±7.8 76.7±4.3 
IC50L-NNA 25.7±10.7 39.8±20.3 
IC50L-NMMA 45.2±23.5 44.3±19.2 
Concentration (μM)
Physarum NOS APhysarum NOS B
KML-arginine 66.7±14.5 26.8±9.7 
EC50 H4-bip 7.9 ±1.2 0.7±0.1 
EC50 NADPH 57.2±6.5 60.3±13.8 
EC50 FAD 3.9±1.6 2.1±1.6 
EC50 FMN 1.7±0.6 0.2±0.2 
IC50L-NIL 81.3±7.8 76.7±4.3 
IC50L-NNA 25.7±10.7 39.8±20.3 
IC50L-NMMA 45.2±23.5 44.3±19.2 

Haem content of homogeneous Physarum NOSs A1 and A1LD

To check whether or not the mutation of a residue in the N-terminus that led to a pronounced decrease in activity also affected haem binding and hence haem content of the active enzyme, we introduced into forms A1 and A1LD (see Figure 3A) a StrepTag, and further purified the tagged proteins by additional gel filtration, collection of the dimeric fractions and final purification using the StrepTag affinity column. This yielded a homogeneous protein (>95% SDS gel silver–stained; Figure 6C) free of E. coli impurities (minor impurities also contain the StrepTag as checked by Western blotting). Pure A1 (three preparations) had a specific activity of 585±144 nmol·mg−1·min−1, and a haem content of 0.41±0.14 haem per subunit, whereas pure A1LD (two preparations) displayed less than one-tenth of the activity (51±10 nmol·mg−1·min−1), but had a haem content (0.44±0.09 per subunit) comparable with A1 (Table 2).

Table 2
Original data for haem determination of pure Physarum NOS preparations

Physarum NOS clones A1 and A1LD containing an N-terminal Strep tag were co-expressed with Physarum calmodulin in E. coli with pETDuet1 and purified by ammonium sulfate fractionation, ADP-Sepharose-affinity chromatography, gel filtration and Strep-tag-affinity chromatography. Haem was determined in purified fractions by reversed-phase HPLC with UV detection. NOS activity was determined by the radiometric citrulline assay. Protein was determined by the Bradford assay using BSA as the standard. MM, calculated molecular mass. See the Experimental section for details of the method.

SampleBatchHaem (μM)Protein (mg/ml)MM (kDa)Haem/subunitNOS activity (nmol·mg−1·min−1)
NOSA1 2.57 1.30 126 0.25 768 
NOSA1 7.47 2.32 126 0.41 417 
NOSA1 4.71 1.02 126 0.58 569 
NOSA1-LD 5.01 1.80 126 0.35 41 
NOSA1-LD 4.16 1.00 126 0.52 61 
SampleBatchHaem (μM)Protein (mg/ml)MM (kDa)Haem/subunitNOS activity (nmol·mg−1·min−1)
NOSA1 2.57 1.30 126 0.25 768 
NOSA1 7.47 2.32 126 0.41 417 
NOSA1 4.71 1.02 126 0.58 569 
NOSA1-LD 5.01 1.80 126 0.35 41 
NOSA1-LD 4.16 1.00 126 0.52 61 

P. polycephalum, a member of the class/superclass Myxogastridae (also termed myxomycetes or true slime molds), is the only so far identified non-animal species known to express NOS with the characteristics of mammalian iNOS [17]. Previously, we showed that this enzyme is induced by starvation of macroplasmodia and plays a crucial role in sporulation [18]. Cloning experiments indicated that Physarum contains two highly related calcium-independent NOSs (forms A and B, 82% amino acid identity) that are expressed in parallel throughout various developmental stages and have a similar molecular mass, as was not only deduced from the cDNA sequences but was underlined by the fact that purified native Physarum NOS was a mixture of both proteins that were not separated by SDS/PAGE [18]. While parallel expression of calcium-dependent and calcium-independent NOS isoforms within a certain cell type is common in animal species, the parallel occurrence of two versions of calcium-independent NOS has not been reported, at least to our knowledge. The previously used Physarum strain was the diploid M3b isolate. Thus it was possible that the two different cDNAs and proteins could stem from two alleles of the same gene, as it was described for the plasmodiumspecific hapP mRNA that encodes proteins of unknown function differing by 9.6% in their predicted amino acid sequence [26]. We therefore checked for expression of Physarum NOS form A and form B in the haploid Physarum strain LU352 [23], the strain now being used for sequencing the Physarum genome (http://www.genome.gov/12511858 and http://genome.wustl.edu/genome.cgi?GENOME=Physarum%20polycephalum). Quantifying Physarum NOS form A and form B expression in starved macroplasmodia of both strains showed comparable mRNA levels. Moreover, both mRNAs were also expressed in LU352 amoebae. Macroplasmodia of diploid strains develop after fusion of genetically different amoebae, but in the apogamic haploid strains like LU352 no such fusion is required [21,30]. Therefore expression of both NOS mRNAs in LU352 amoebae underlined the assumption that the two Physarum NOS forms stem from two genes rather than from two alleles of the same gene.

Studying the organization of these two genes revealed the high similarity of the overall structure of the Physarum NOS genes to those of NOS genes from higher animals including mammals. Intron size of Physarum NOSs is smaller than that of animal NOSs, which is in agreement with the smaller genome size of Physarum (300 Mb [30a,30b]) as compared with human (3038 Mb) or chicken (1200 Mb; http://www.ncbi.nlm.nih.gov/genomes/leuks.cgi) [31]. None of the intron positions is conserved between Physarum and animal NOS genes, which is in line with the idea that introns are transposable elements that are acquired or also deleted during evolution (reviewed in [32]). Transposition of introns occurs very infrequently during evolution; for example, most orthologous genes between mouse and human, species that diverged at least 50 million years ago, have the same intron positions [32]. This gives some idea of the evolutionary distance between Physarum and higher animals and their early divergence from a common ancestor (see below) where already some ancestral NOS had evolved. Once the final Physarum genome sequence is available, these speculations can be studied in more detail and it will also be possible to clarify whether the two Physarum NOS genes stem from single gene duplication or other evolutionary events.

Some further interesting aspects with regard to evolution become evident here. According to molecular phylogenetic data, Physarum is most closely related to the cellular slime moulds, the Dictyostelidae [33,34]. Together with other amoebae and slime mould classes, they are grouped as Amoebozoa, one of eight major groups composing the ‘crown’ of the phylogenetic tree of eukaryotes [33,35], thus indicating that these organisms cannot be considered as ‘low’ stages on the path to higher animals and fungi but have developed independently from a more ancient, not yet defined, root [35]. However, a more refined analysis of evolutionary relationships between Physarum and Dictyostelium, another model organism for cell biology, has to await the final Physarum genome. In any case, our results on NOS as well as previous observations concerning pteridine biosynthesis (see below) underline the diversity of Physarum and Dictyostelium despite the striking similarities of their life cycle, the occurrence of mobile and stationary developmental stages, their habitat and their way of feeding as well as the fact that their protein sequences usually group on a common branch in phylogenetic analysis. First of all, Dictyostelium discoideum does not contain any gene similar to full-length NOS (as checked by BLAST-P search of the Dictyostelium genome with the Physarum NOSs; results not shown), whereas Physarum contains two genes with exon numbers comparable with those of higher animals (see Figure 2) and an even intron distribution characteristic of higher eukaryotes [36]. Secondly, Dictyostelium does not produce H4-bip like Physarum [17] but its stereoisomer tetrahydrodictyopterin [37], a pterin that cannot serve as a cofactor for mammalian NOS [38] or Physarum NOS (E. R. Werner, G. Golderer, P. Gröbner and G. Werner-Felmayer, unpublished work). Moreover, the genomic structures of GTP cyclohydrolase I, the first enzyme of H4-bip biosynthesis, differ significantly between the two species [39]: in accordance with its comparatively small genome of 34 Mb (http://www.ncbi.nlm.nih.gov/genomes/leuks.cgi), Dictyostelium has only one 109 bp intron [40], whose location is conserved in Physarum, whereas the Physarum gene consists of 7 exons and thus resembles animal GTP cyclohydrolase I genes with some intron positions conserved in Drosophila or human. In addition, alternative splicing of GTP cyclohydrolase I occurs in Physarum at a similar position to human [39]. Taken together, these findings suggest that Physarum may be more closely related to animals than Dictyostelium with respect of H4-bip and nitric oxide synthesis, and it underlines the value of Physarum as a model system for studying certain aspects of animal cell biology. In line with our observations, the recent analysis of Physarum expressed sequence tags [41] has revealed 895 genes of Physarum with similarities in other databases but not in the Dictyostelium protein database. In addition, approx. 9% of all detected genes are alternatively spliced in Physarum, whereas only approx. 0.2% of genes are alternatively spliced in Dictyostelium [41].

In the course of setting up recombinant expression of both Physarum NOSs in bacteria, we found that the previously predicted reading frame [18] was too short at its N-terminus. We concluded this not only from lack of activity of His-tagged expression constructs, but also from the smaller size of the recombinant proteins as compared with the native NOS purified from Physarum microplasmodia (Figure 3C). Addition of only a few amino acids (13 for NOS form A and four for NOS form B) to the N-terminus as deduced from the cDNAs led to expression of functional proteins. Producing 5′-extended cDNA sequences for both NOSs and checking for other possible translation starts revealed that only NOS form B had an upstream methionine within reasonable distance, whereas for NOS form A we found a stop codon and no additional AUG codons between this stop and the previously predicted start. Thus at least NOS form A appears to use a non-AUG start codon [29] and this may also be the case for NOS form B. The molecular masses of the predicted NOSs starting from these alternative start codons (both encoding isoleucine) compare much better with native NOS than those of the proteins starting with the first encoded AUG codon. It should be noted that none of the possible AUG codons have an optimal context, whereas the predicted alternative non-AUG codons, particularly of NOS form B, fit much better to this rule [29]. Future studies will clarify the exact start of Physarum NOS proteins, but at least NOS form A and presumably also form B appear to belong to the growing group of mRNAs with non-AUG translation initiation [29]. Sometimes, as was shown for the mammalian translational regulator EIF4G2 (eukaryotic translation initiation factor 4γ2) [28], usage of a non-AUG translation initiation site is even conserved evolutionarily. However, there is no report that any of the animal NOS mRNAs starts translation at a non-AUG codon.

Although the two isoforms of Physarum NOS are highly similar, we found distinct differences in the enzymology of the recombinant proteins expressed in E. coli. Physarum NOS form B has a significantly higher affinity towards L-arginine, H4-bip and FMN than Physarum NOS form A. The purified recombinant proteins of form A and B contained about the same amount of dimer and monomer, and were partly decoupled in that they consumed more than 5 NADPH molecules per nitric oxide molecule formed rather than 1.5 NADPH molecules per nitric oxide molecule for the fully coupled mammalian enzyme [11].

We were surprised to find the requirement of N-terminal amino acid residues upstream of the first AUG codon since this is a region with low homology with other NOSs, and a region that iNOS oxygenase domains apparently do not require for function [42]. In iNOS oxygenase domain crystals, this region is disordered [43] so that no structural information on this region is available to date. After we had determined the minimum amount of residues required to yield a functional recombinant protein, we mutated the amino acid on the truncation of which the activity was lost, a leucine residue, in a larger construct to investigate whether we simply required a certain length of the protein, or whether specific residues were needed. Activity was lost on mutation of this leucine residue to aspartate or glycine, without affecting the monomer/dimer ratio and the haem content of homogeneously purified proteins. We assume that specific residues in this region are possibly required for interactions between oxygenase and reductase domains in the dimer of the enzyme, for which hydrophobic forces might be crucial.

We thank Renate Kaus, Petra Loitzl and Nina Madl for excellent technical assistance. We also thank Wolfgang Marwan (University of Magdeburg, Germany) for providing the LU352 strain and Jonatha Gott (Case Western Reserve University, Cleveland, OH, U.S.A.) for the Physarum genomic library.

NOS

nitric oxide synthase

iNOS

inducible NOS

TB

Terrific broth

DTE

dithioerythritol

H4-bip

6R-5,6,7,8-tetrahydro-L-biopterin

LB

Luria–Bertani

RACE

rapid amplification of cDNA ends

This work was supported by the Austrian Funds ‘Zur Förderung der wissenschaftlichen Forschung’, [grant numbers P15538 (to G. G.), P16059 (to G. W.-F.)], and by the Tiroler Wissenschaftsfonds [grant number UNI-0404-245 (to G. W.-F.)].

1
Forstermann
 
U.
Boissel
 
J. P.
Kleinert
 
H.
 
Expressional control of the ‘constitutive’ isoforms of nitric oxide synthase (NOS I and NOS III)
FASEB J.
1998
, vol. 
12
 (pg. 
773
-
790
)
2
Eissa
 
N. T.
Yuan
 
J. W.
Haggerty
 
C. M.
Choo
 
E. K.
Palmer
 
C. D.
Moss
 
J.
 
Cloning and characterization of human inducible nitric oxide synthase splice variants: a domain, encoded by exons 8 and 9, is critical for dimerization
Proc. Natl. Acad. Sci. U.S.A.
1998
, vol. 
95
 (pg. 
7625
-
7630
)
3
Tiscornia
 
A. C.
Cayota
 
A.
Landoni
 
A. I.
Brito
 
C.
Oppezzo
 
P.
Vuillier
 
F.
Robello
 
C.
Dighiero
 
G.
Gabus
 
R.
Pritsch
 
O.
 
Post-transcriptional regulation of inducible nitric oxide synthase in chronic lymphocytic leukemia B cells in pro- and antiapoptotic culture conditions
Leukemia
2004
, vol. 
18
 (pg. 
48
-
56
)
4
Bros
 
M.
Boissel
 
J. P.
Godtel-Armbrust
 
U.
Forstermann
 
U.
 
Transcription of human neuronal nitric oxide synthase mRNAs derived from different first exons is partly controlled by exon 1-specific promoter sequences
Genomics
2006
, vol. 
87
 (pg. 
463
-
473
)
5
Guix
 
F. X.
Uribesalgo
 
I.
Coma
 
M.
Munoz
 
F. J.
 
The physiology and pathophysiology of nitric oxide in the brain
Prog. Neurobiol.
2005
, vol. 
76
 (pg. 
126
-
152
)
6
Mungrue
 
I. N.
Bredt
 
D. S.
 
nNOS at a glance: implications for brain and brawn
J. Cell Sci.
2004
, vol. 
117
 (pg. 
2627
-
2629
)
7
Bronte
 
V.
Zanovello
 
P.
 
Regulation of immune responses by L-arginine metabolism
Nat. Rev. Immunol.
2005
, vol. 
5
 (pg. 
641
-
654
)
8
Belge
 
C.
Massion
 
P. B.
Pelat
 
M.
Balligand
 
J. L.
 
Nitric oxide and the heart: update on new paradigms
Ann. N.Y. Acad. Sci.
2005
, vol. 
1047
 (pg. 
173
-
182
)
9
Danson
 
E. J.
Paterson
 
D. J.
 
Cardiac neurobiology of nitric oxide synthases
Ann. N.Y. Acad. Sci.
2005
, vol. 
1047
 (pg. 
183
-
196
)
10
Griffith
 
O. W.
Stuehr
 
D. J.
 
Nitric oxide synthases: properties and catalytic mechanism
Annu. Rev. Physiol.
1995
, vol. 
57
 (pg. 
707
-
736
)
11
Gorren
 
A. C.
Mayer
 
B.
 
Nitric-oxide synthase: a cytochrome P450 family foster child
Biochim. Biophys. Acta
2007
, vol. 
1770
 (pg. 
432
-
445
)
12
Stuehr
 
D. J.
Santolini
 
J.
Wang
 
Z. Q.
Wei
 
C. C.
Adak
 
S.
 
Update on mechanism and catalytic regulation in the NO synthases
J. Biol. Chem.
2004
, vol. 
279
 (pg. 
36167
-
36170
)
13
Panda
 
K.
Rosenfeld
 
R. J.
Ghosh
 
S.
Meade
 
A. L.
Getzoff
 
E. D.
Stuehr
 
D. J.
 
Distinct dimer interaction and regulation in nitric-oxide synthase types I, II, and III
J. Biol. Chem.
2002
, vol. 
277
 (pg. 
31020
-
31030
)
14
Wei
 
C. C.
Crane
 
B. R.
Stuehr
 
D. J.
 
Tetrahydrobiopterin radical enzymology
Chem. Rev.
2003
, vol. 
103
 (pg. 
2365
-
2383
)
15
Werner
 
E. R.
Gorren
 
A. C.
Heller
 
R.
Werner-Felmayer
 
G.
Mayer
 
B.
 
Tetrahydrobiopterin and nitric oxide: mechanistic and pharmacological aspects
Exp. Biol. Med.
2003
, vol. 
228
 (pg. 
1291
-
1302
)
16
Gorren
 
A. C.
Sorlie
 
M.
Andersson
 
K. K.
Marchal
 
S.
Lange
 
R.
Mayer
 
B.
 
Tetrahydrobiopterin as combined electron/proton donor in nitric oxide biosynthesis: cryogenic UV–Vis and EPR detection of reaction intermediates
Methods Enzymol.
2005
, vol. 
396
 (pg. 
456
-
466
)
17
Werner-Felmayer
 
G.
Golderer
 
G.
Werner
 
E. R.
Grobner
 
P.
Wachter
 
H.
 
Pteridine biosynthesis and nitric oxide synthase in Physarum polycephalum
Biochem. J.
1994
, vol. 
304
 (pg. 
105
-
111
)
18
Golderer
 
G.
Werner
 
E. R.
Leitner
 
S.
Grobner
 
P.
Werner-Felmayer
 
G.
 
Nitric oxide synthase is induced in sporulation of Physarum polycephalum
Genes Dev.
2001
, vol. 
15
 (pg. 
1299
-
1309
)
19
Buddha
 
M. R.
Tao
 
T.
Parry
 
R. J.
Crane
 
B. R.
 
Regioselective nitration of tryptophan by a complex between bacterial nitric-oxide synthase and tryptophanyl-tRNA synthetase
J. Biol. Chem.
2004
, vol. 
279
 (pg. 
49567
-
49570
)
20
Daniel
 
J.
Baldwin
 
H. H.
 
Prescott
 
D. M.
 
Methods of culture for plasmodial myxomycetes
Methods of Cell Physiology, Volume 1
1964
New York
Academic Press
(pg. 
9
-
41
)
21
Dee
 
J.
Foxon
 
J. L.
Anderson
 
R. W.
 
Growth, development and genetic-characteristics of Physarum polycephalum amebas able to grow in liquid, axenic medium
J. Gen. Microbiol.
1989
, vol. 
135
 (pg. 
1567
-
1588
)
22
Schmidt
 
T. G.
Skerra
 
A.
 
The Strep-tag system for one-step purification and high-affinity detection or capturing of proteins
Nat. Protocol
2007
, vol. 
2
 (pg. 
1528
-
1535
)
23
Barroso
 
J. F.
Elholm
 
M.
Flatmark
 
T.
 
Tight binding of deoxyribonucleotide triphosphates to human thymidine kinase 2 expressed in Escherichia coli. Purification and partial characterization of its dimeric and tetrameric forms
Biochemistry
2003
, vol. 
42
 (pg. 
15158
-
15169
)
24
Gross
 
S. S.
 
Microtiter plate assay for determining kinetics of nitric oxide synthesis
Methods Enzymol.
1996
, vol. 
268
 (pg. 
159
-
168
)
25
Bonkovsky
 
H. L.
Wood
 
S. G.
Howell
 
S. K.
Sinclair
 
P. R.
Lincoln
 
B.
Healey
 
J. F.
Sinclair
 
J. F.
 
High-performance liquid chromatographic separation and quantitation of tetrapyrroles from biological materials
Anal. Biochem.
1986
, vol. 
155
 (pg. 
56
-
64
)
26
Lepine
 
G.
Laroche
 
A.
Lemieux
 
G.
Pallotta
 
D.
 
The two alleles of the hapP gene in Physarum polycephalum code for different proteins
Biochim. Biophys. Acta
1995
, vol. 
1264
 (pg. 
271
-
274
)
27
Wu
 
C.
Zhang
 
J.
Abu-Soud
 
H.
Ghosh
 
D. K.
Stuehr
 
D. J.
 
High-level expression of mouse inducible nitric oxide synthase in Escherichia coli requires coexpression with calmodulin
Biochem. Biophys. Res. Commun.
1996
, vol. 
222
 (pg. 
439
-
444
)
28
Takahashi
 
K.
Maruyama
 
M.
Tokuzawa
 
Y.
Murakami
 
M.
Oda
 
Y.
Yoshikane
 
N.
Makabe
 
K. W.
Ichisaka
 
T.
Yamanaka
 
S.
 
Evolutionarily conserved non-AUG translation initiation in NAT1/p97/DAP5 (EIF4G2)
Genomics
2005
, vol. 
85
 (pg. 
360
-
371
)
29
Kozak
 
M.
 
Regulation of translation via mRNA structure in prokaryotes and eukaryotes
Gene
2005
, vol. 
361
 (pg. 
13
-
37
)
30
Bailey
 
J.
Anderson
 
R. W.
Dee
 
J.
 
Growth and development in relation to the cell cycle in Physarum polycephalum
Protoplasma
1987
, vol. 
141
 (pg. 
101
-
111
)
30a
Mohberg
 
J.
Babcock
 
K. L.
Haugli
 
F. B.
Rusch
 
H. P.
 
Nuclear DNA content and chromosome numbers in the myxomycete Physarum polycephalum
Dev. Biol.
1973
, vol. 
34
 (pg. 
228
-
245
)
30b
Kubbies
 
M.
Pierron
 
G.
 
Mitotic cell cycle control in Physarum. Unprecedented insights via flow-cytometry
Exp. Cell Res.
1983
, vol. 
149
 (pg. 
57
-
67
)
31
Charlesworth
 
B.
Barton
 
N.
 
Genome size: does bigger mean worse?
Curr. Biol.
2004
, vol. 
14
 (pg. 
R233
-
R235
)
32
Fedorova
 
L.
Fedorov
 
A.
 
Introns in gene evolution
Genetica
2003
, vol. 
118
 (pg. 
123
-
131
)
33
Baldauf
 
S. L.
Doolittle
 
W. F.
 
Origin and evolution of the slime molds (Mycetozoa)
Proc. Natl. Acad. Sci. U.S.A.
1997
, vol. 
94
 (pg. 
12007
-
12012
)
34
Baldauf
 
S. L.
Roger
 
A. J.
Wenk-Siefert
 
I.
Doolittle
 
W. F.
 
A kingdom-level phylogeny of eukaryotes based on combined protein data
Science
2000
, vol. 
290
 (pg. 
972
-
977
)
35
Baldauf
 
S. L.
 
The deep roots of eukaryotes
Science
2003
, vol. 
300
 (pg. 
1703
-
1706
)
36
Mourier
 
T.
Jeffares
 
D. C.
 
Eukaryotic intron loss
Science
2003
, vol. 
300
 pg. 
1393
 
37
Klein
 
R.
Thiery
 
R.
Tatischeff
 
I.
 
Dictyopterin, 6-(D-threo-1,2-dihydroxypropyl)-pterin, a new natural isomer of L-biopterin. Isolation from vegetative cells of Dictyostelium discoideum and identification
Eur. J. Biochem.
1990
, vol. 
187
 (pg. 
665
-
669
)
38
Werner
 
E. R.
Pitters
 
E.
Schmidt
 
K.
Wachter
 
H.
Werner-Felmayer
 
G.
Mayer
 
B.
 
Identification of the 4-amino analogue of tetrahydrobiopterin as a dihydropteridine reductase inhibitor and a potent pteridine antagonist of rat neuronal nitric oxide synthase
Biochem. J.
1996
, vol. 
320
 (pg. 
193
-
196
)
39
Golderer
 
G.
Werner
 
E. R.
Heufler
 
C.
Strohmaier
 
W.
Grobner
 
P.
Werner-Felmayer
 
G.
 
GTP cyclohydrolase I mRNA: novel splice variants in the slime mould Physarum polycephalum and in human monocytes (THP-1) indicate conservation of mRNA processing
Biochem. J.
2001
, vol. 
355
 (pg. 
499
-
507
)
40
Witter
 
K.
Cahill
 
D. J.
Werner
 
T.
Ziegler
 
I.
Rodl
 
W.
Bacher
 
A.
Gutlich
 
M.
 
Molecular cloning of a cDNA coding for GTP cyclohydrolase I from Dictyostelium discoideum
Biochem. J.
1996
, vol. 
319
 (pg. 
27
-
32
)
41
Gloeckner
 
G.
Golderer
 
G.
Werner-Felmayer
 
G.
Meyer
 
S.
Marwan
 
W.
 
A first glimpse at the transcriptome of Physarum polycephalum
BMC Genomics
2008
, vol. 
9
 pg. 
6
 
42
Ghosh
 
D. K.
Crane
 
B. R.
Ghosh
 
S.
Wolan
 
D.
Gachhui
 
R.
Crooks
 
C.
Presta
 
A.
Tainer
 
J. A.
Getzoff
 
E. D.
Stuehr
 
D. J.
 
Inducible nitric oxide synthase: role of the N-terminal beta-hairpin hook and pterin-binding segment in dimerization and tetrahydrobiopterin interaction
EMBO J.
1999
, vol. 
18
 (pg. 
6260
-
6270
)
43
Crane
 
B. R.
Arvai
 
A. S.
Ghosh
 
D. K.
Wu
 
C.
Getzoff
 
E. D.
Stuehr
 
D. J.
Tainer
 
J. A.
 
Structure of nitric oxide synthase oxygenase dimer with pterin and substrate
Science
1998
, vol. 
279
 (pg. 
2121
-
2126
)
44
Bredt
 
D. S.
Hwang
 
P. M.
Glatt
 
C. E.
Lowenstein
 
C.
Reed
 
R. R.
Snyder
 
S. H.
 
Cloned and expressed nitric oxide synthase structurally resembles cytochrome P-450 reductase
Nature
1991
, vol. 
351
 (pg. 
714
-
718
)
45
Raman
 
C. S.
Li
 
H.
Martasek
 
P.
Kral
 
V.
Masters
 
B. S.
Poulos
 
T. L.
 
Crystal structure of constitutive endothelial nitric oxide synthase: a paradigm for pterin function involving a novel metal center
Cell
1998
, vol. 
95
 (pg. 
939
-
950
)
46
Fischmann
 
T. O.
Hruza
 
A.
Niu
 
X. D.
Fossetta
 
J. D.
Lunn
 
C. A.
Dolphin
 
E.
Prongay
 
A. J.
Reichert
 
P.
Lundell
 
D. J.
Narula
 
S. K.
Weber
 
P. C.
 
Structural characterization of nitric oxide synthase isoforms reveals striking active-site conservation
Nat. Struct. Biol.
1999
, vol. 
6
 (pg. 
233
-
242
)
47
Garcia-Cardena
 
G.
Martasek
 
P.
Masters
 
B. S.
Skidd
 
P. M.
Couet
 
J.
Li
 
S.
Lisanti
 
M. P.
Sessa
 
W. C.
 
Dissecting the interaction between nitric oxide synthase (NOS) and caveolin. Functional significance of the NOS caveolin binding domain in vivo
J. Biol. Chem.
1997
, vol. 
272
 (pg. 
25437
-
25440
)

Author notes

1

Present address: Institute of Veterinary Biochemistry and Molecular Biology, University of Zürich, Winterthurerstr. 190, CH-8057 Zürich, Switzerland.

The nucleotide sequence data reported will appear in GenBank®, EMBL, DDBJ and GSDB Nucleotide Sequence Databases under the accession numbers DQ835529, DQ845107, DQ835525, DQ835527, DQ835526 and DQ825528.

This is an open-access article distributed under the terms of the Creative Commons Attribution Non-Commercial Licence (http://creativecommons.org/licenses/by-nc/2.5/) which permits unrestricted non-commercial use, distribution and reproduction in any medium, provided the original work is properly cited.