21.08.2013 Views

Analysis of a ferric leghemoglobin reductase from cowpea (Vigna ...

Analysis of a ferric leghemoglobin reductase from cowpea (Vigna ...

Analysis of a ferric leghemoglobin reductase from cowpea (Vigna ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Plant Science 154 (2000) 161–170<br />

<strong>Analysis</strong> <strong>of</strong> a <strong>ferric</strong> <strong>leghemoglobin</strong> <strong>reductase</strong> <strong>from</strong> <strong>cowpea</strong> (<strong>Vigna</strong><br />

unguiculata) root nodules<br />

Peng Luan a , Elena Aréchaga-Ocampo b , Gautam Sarath a , Raúl Arredondo-Peter b ,<br />

Robert V. Klucas a, *<br />

a Department <strong>of</strong> Biochemistry, The Beadle Center, Uniersity <strong>of</strong> Nebraska-Lincoln, Lincoln, NE 68588-0664, USA<br />

b Centro de Inestigación sobre Fijación de Nitrógeno, Uniersidad Nacional Autónoma de México, Apartado Postal 565-A, 62210 Cuernaaca,<br />

Morelos, Mexico<br />

Abstract<br />

Received 2 September 1999; received in revised form 9 December 1999; accepted 15 December 1999<br />

www.elsevier.com/locate/plantsci<br />

Ferric <strong>leghemoglobin</strong> <strong>reductase</strong> (FLbR), an enzyme reducing <strong>ferric</strong> <strong>leghemoglobin</strong> (Lb) to ferrous Lb, was purified <strong>from</strong> <strong>cowpea</strong><br />

(<strong>Vigna</strong> unguiculata) root nodules by sequential chromatography on hydroxylapatite followed by Mono-Q HR5/5 FPLC and<br />

Sephacryl S-200 gel filtration. The purified <strong>cowpea</strong> FLbR had a specific activity <strong>of</strong> 216 nmol Lb 2+ O 2 formed min −1 mg −1 <strong>of</strong><br />

enzyme for <strong>cowpea</strong> Lb 3+ and a specific activity <strong>of</strong> 184 nmol Lb 2+ O 2 formed min −1 mg −1 <strong>of</strong> enzyme for soybean Lb 3+ .A<br />

cDNA clone <strong>of</strong> <strong>cowpea</strong> FLbR was obtained by screening a <strong>cowpea</strong> root nodule cDNA library. The nucleotide sequence <strong>of</strong> <strong>cowpea</strong><br />

FLbR cDNA exhibited about 88% similarity with soybean (Glycine max) FLbR and 85% with pea (Pisum satium) dihydrolipoamide<br />

dehydrogenase (DLDH, EC 1.8.1.4) cDNAs. Conserved regions for the FAD-binding site, NAD(P)H-binding site,<br />

and disulfide active site were identified among the deduced amino acid sequences <strong>of</strong> <strong>cowpea</strong> FLbR, soybean FLbR, pea DLDH<br />

and other enzymes in the family <strong>of</strong> the pyridine nucleotide-disulfide oxido-<strong>reductase</strong>s. © 2000 Published by Elsevier Science<br />

Ireland Ltd. All rights reserved.<br />

Keywords: Cowpea; Dihyrodrolipoamide dehydrogenase; Ferric <strong>leghemoglobin</strong> <strong>reductase</strong>; Leghemoglobin; Symbiotic nitrogen fixation; <strong>Vigna</strong><br />

unguiculata<br />

1. Introduction<br />

Leghemoglobins (Lbs) are important nodule<br />

proteins that reversibly bind O 2 and facilitate its<br />

diffusion to the N 2-fixing bacteroids in root nodules.<br />

This provides a flux <strong>of</strong> O 2 for rhizobial<br />

respiration, while maintaining O 2 at a concentration<br />

that does not inactivate the nitrogenase complex<br />

[1]. Lb can exist in several different oxidation<br />

states: ferrous (Lb 2+ ), <strong>ferric</strong> (Lb 3+ ) or ferryl form<br />

(Lb 4+ ), but only Lb 2+ is functional. Slight<br />

changes in the physiology <strong>of</strong> nodules, such as the<br />

Abbreiations: DLDH, dihydrolipoamide dehydrogenase; FLbR,<br />

<strong>ferric</strong> <strong>leghemoglobin</strong> <strong>reductase</strong>; Lb, <strong>leghemoglobin</strong>.<br />

* Corresponding author. Tel.: +1-402-4722932; fax: +1-402-<br />

4727842.<br />

E-mail address: rklucas@unlnotes.unl.edu (R.V. Klucas)<br />

0168-9452/00/$ - see front matter © 2000 Published by Elsevier Science Ireland Ltd. All rights reserved.<br />

PII: S0168-9452(99)00272-1<br />

presence <strong>of</strong> some metal ions, chelators, and toxic<br />

metabolites (nitrite, superoxide radical, peroxides),<br />

may cause the oxidation <strong>of</strong> functional Lb 2+ into<br />

the nonfunctional Lb 3+ and Lb 4+ [2]. Mechanisms<br />

must therefore exist in legume plants for<br />

maintaining Lb in its functional Lb 2+ state.<br />

Enzymatic reduction <strong>of</strong> Lb 3+ to Lb 2+ has been<br />

hypothesized to exist in legume nodules [3]. A<br />

protein that reduces Lb 3+ to Lb 2+ was purified<br />

<strong>from</strong> lupin nodules. It had a molecular weight <strong>of</strong><br />

60 kDa, contained FAD as a c<strong>of</strong>actor, used<br />

NADH as the electron donor, methylene blue as<br />

the electron carrier, and had K m values <strong>of</strong> 8.7 M<br />

for NADH and 10 M for lupin Lb 3+ [4,5].<br />

Another protein, FLbR, was purified to homogeneity<br />

<strong>from</strong> soybean root nodules and further<br />

characterized in our laboratory [6,7]. The purified


162<br />

soybean FLbR was a homodimer with a molecular<br />

weight <strong>of</strong> 110 kDa, contained FAD as a<br />

prosthetic group, and used NAD(P)H as the<br />

electron donor to reduce Lb 3+ . It also exhibited<br />

diaphorase activity [6]. The enzyme required O 2<br />

at the micromolar levels for the reduction <strong>of</strong><br />

Lb 3+ to Lb 2+ in vitro, had K m values <strong>of</strong> 7 M<br />

for NADH, 9.5 M for soybean Lb 3+ , and a<br />

V max value for soybean Lb 3+ reduction <strong>of</strong> 499<br />

nmol Lb 2+ O 2 formed min −1 mg −1 [7]. A cDNA<br />

encoding soybean FLbR was cloned and sequenced<br />

[8] and subsequently overexpressed in<br />

Escherichia coli [9]. Based on sequence homology,<br />

soybean FLbR was shown to be related to<br />

a family <strong>of</strong> pyridine nucleotide-disulfide oxido-<strong>reductase</strong>s,<br />

especially the DLDHs <strong>from</strong> various<br />

sources [9].<br />

The only FLbR that had been characterized to<br />

date is <strong>from</strong> soybean root nodules. To investigate<br />

if this enzyme is present in other legumes<br />

nodules, we used <strong>cowpea</strong> (<strong>Vigna</strong> unguiculata)<br />

root nodules for our study. In this work we describe<br />

(1) the identification and purification <strong>of</strong><br />

FLbR <strong>from</strong> <strong>cowpea</strong> root nodules; (2) some <strong>of</strong><br />

the important physical and enzymatic properties<br />

<strong>of</strong> <strong>cowpea</strong> FLbR; (3) the sequences <strong>of</strong> the<br />

cDNA encoding <strong>cowpea</strong> FLbR; and (4) the comparison<br />

<strong>of</strong> deduced <strong>cowpea</strong> FLbR amino acids<br />

sequence with soybean FLbR and other similar<br />

proteins. The results revealed that the <strong>cowpea</strong><br />

FLbR is very similar to the soybean enzyme,<br />

and indeed FLbR may be common to all<br />

legumes root nodules.<br />

2. Materials and methods<br />

2.1. Chemicals<br />

Chemicals were reagent or molecular biology<br />

grade. Bio-Gel-hydroxylapatite, Bio-Gel P6-DG,<br />

silver staining kit, and protein concentration assay<br />

kit were purchased <strong>from</strong> Bio Rad. Prepacked<br />

Mono-Q HR5/5 column, Sephacryl S-200<br />

Super Fine (SF), Sephadex G-25 and G-75 were<br />

<strong>from</strong> Pharmacia. Agarose, bacteria broth, Taq<br />

DNA polymerase, PCR reagents and DNA ligation<br />

reagents were <strong>from</strong> Gibco-BRL. Hybridization<br />

materials were <strong>from</strong> Boehringer Mannheim.<br />

Other chemicals were <strong>from</strong> Sigma and Fisher unless<br />

noted.<br />

P. Luan et al. / Plant Science 154 (2000) 161–170<br />

2.2. Purification <strong>of</strong> <strong>cowpea</strong> FLbR <strong>from</strong> root<br />

nodules<br />

Germinated <strong>cowpea</strong> seeds (V. unguiculata, cv.<br />

Blackeye peas, 137-California No. 5) were<br />

inoculated with Bradyrhizobium japonicum USDA<br />

3456 before planting. The bacteria and plants were<br />

grown as described by Becana et al. [10]. Cowpea<br />

root nodules were harvested <strong>from</strong> 5 weeks old<br />

plants, and stored at −90°C. All purification<br />

steps were carried out at 4°C, essentially following<br />

the procedure <strong>of</strong> Ji et al. [7]. Active fractions were<br />

collected, pooled, made to a final concentration <strong>of</strong><br />

10% with glycerol and stored at −90°C.<br />

Homogeneity <strong>of</strong> the sample after each purification<br />

step was analyzed by 10% SDS-PAGE stained<br />

using the Bio-Rad Silver-Stain Kit. Protein<br />

concentration was determined using Bio Rad<br />

protein micro-assay procedure and bovine serum<br />

albumin as a standard.<br />

2.3. Isolation <strong>of</strong> <strong>cowpea</strong> Lb and soybean Lb<br />

Isolation and oxidization <strong>of</strong> <strong>cowpea</strong> and soybean<br />

Lb were as described earlier [10,11]. Lb<br />

was stored as Lb 3+ at −90°C until use.<br />

2.4. FLbR actiity and diaphorase actiity assay<br />

FLbR activity was assayed according to the<br />

procedure used by Ji et al. [7] on a Milton Roy<br />

Spectronic 3000 Array spectrophotometer<br />

equipped with kinetic acquisition s<strong>of</strong>tware. The<br />

absorbance change was converted to Lb 2+ O 2<br />

formation rate using the n <strong>of</strong> 10.2 mM −1 cm −1<br />

(Lb 2+ O 2 minus Lb 3+ ) at 574 nm. FLbR specific<br />

activity was expressed as nmol Lb 2+ O 2 formed<br />

min −1 mg −1 <strong>of</strong> protein. Diaphorase activity was<br />

assayed using DCPIP as described by Ji et al. [7].<br />

Diaphorase specific activity was expressed as nmol<br />

DCPIP reduced min −1 mg −1 <strong>of</strong> protein.<br />

2.5. N-Terminal sequencing<br />

Approximately 20 pmol (2 g) <strong>of</strong> purified<br />

<strong>cowpea</strong> FLbR was separated by 10% SDS-PAGE<br />

and then transferred to a polyvinylidene difluoride<br />

membrane (Millipore). The membrane was stained<br />

with Amido Black solution to visualize proteins.<br />

The band corresponding to the FLbR was excised<br />

<strong>from</strong> the membrane and sequenced on an ABI


Instruments Procise 494 Protein Sequencer at the<br />

Protein Core Facility at University <strong>of</strong> Nebraska-<br />

Lincoln.<br />

2.6. Characterization <strong>of</strong> <strong>cowpea</strong> FLbR<br />

Molecular weight <strong>of</strong> native FLbR was determined<br />

using a gel filtration method [12] on a<br />

Sephacryl S-200 SF column. The <strong>cowpea</strong> FLbR<br />

activity was assayed in different pH buffers to<br />

determine its optimal pH, using <strong>cowpea</strong> Lb 3+ as the<br />

substrate. The buffers used were: 50 mM potassium<br />

phosphate at pH 3.7, 4.8; 50 mM MES at pH 5.5,<br />

6.0, 6.5, 7.0; and 50 mM Tris–HCl at pH 7.5, 8.5.<br />

Reduction <strong>of</strong> <strong>cowpea</strong> and soybean Lb 3+ by<br />

<strong>cowpea</strong> FLbR was followed spectrophotometrically<br />

using an assay mixture that contained 50 mM<br />

potassium phosphate, pH 6.5, 2 g <strong>cowpea</strong> FLbR,<br />

500 M NADH, and various concentrations <strong>of</strong><br />

<strong>cowpea</strong> or soybean Lb 3+ (0–50 M) in a final<br />

volume <strong>of</strong> 1 ml. K m and V max values for <strong>cowpea</strong> or<br />

soybean Lb 3+ were determined by fitting the data<br />

to the Michaelis–Menten equation or Lineweaver–<br />

Burk plot using commercial s<strong>of</strong>tware (Grafit version<br />

2.0). The K m value for NADH was determined<br />

as above using reaction mixture that contained 50<br />

mM potassium phosphate, pH 6.5, 50 M <strong>cowpea</strong><br />

Lb, 2 g <strong>cowpea</strong> FLbR, and various concentrations<br />

<strong>of</strong> NADH (0–500 M).<br />

2.7. Construction <strong>of</strong> a <strong>cowpea</strong> root nodule cDNA<br />

library<br />

Poly(A + ) mRNAs were isolated <strong>from</strong> 0.5 g <strong>of</strong><br />

frozen <strong>cowpea</strong> root nodules, and used as a template<br />

to generate cDNAs. The cDNAs were ligated with<br />

a EcoRI/NotI adapter and subsequently ligated to<br />

the gt11 arms following the supplier’s procedures<br />

(Pharmacia). The cDNA-gt11 constructs were<br />

packaged into commercial available phage particles<br />

and then amplified using E. coli strain Y1090<br />

following the suppliers procedures (Pharmacia<br />

Ready-To-Go Lambda Packaging Kit). The cDNA<br />

library was divided into 2-ml aliquots and stored at<br />

−90°C.<br />

2.8. Obtaining the cDNA sequence <strong>of</strong> <strong>cowpea</strong><br />

FLbR<br />

Two primers, named FLbR Fwd (5-<br />

AAATCTCTGTAGACACCA-3) and FLbR Rvs<br />

P. Luan et al. / Plant Science 154 (2000) 161–170 163<br />

(5-GCCTTAGCTCTGCTATTA-3), were designed<br />

based on the soybean FLbR cDNA sequence<br />

(positions 485–502 and 1272–1289, respectively,<br />

<strong>from</strong> Ji et al. [8]). They were used for the amplification<br />

<strong>of</strong> an internal FLbR fragment by PCR at high<br />

stringency (annealing temperature <strong>of</strong> 55°C for 1<br />

min) using aliquots <strong>of</strong> a <strong>cowpea</strong> root nodule cDNA<br />

library (above) as template. PCR products were<br />

resolved in a 1.2% agarose gel, the band <strong>of</strong> the<br />

expected size (800 bp) was cut out and extracted<br />

in 10 l <strong>of</strong> sterile water using a Gene Clean II Kit<br />

(Bio 101) following the supplier’s manual. The<br />

extracted DNA was cloned in the pCR2.1 vector<br />

(Invitrogen) and sequenced (see below), and then<br />

used as a probe for screening the cDNA library<br />

<strong>from</strong> <strong>cowpea</strong> nodules as described by Sambrook et<br />

al. [13]. Positive plaques were picked and eluted into<br />

50 l <strong>of</strong> SM solution [13].<br />

Two primers, named Forward and Reverse,<br />

were used for the amplification <strong>of</strong> inserts <strong>from</strong> the<br />

positive plaques essentially as described by Arredondo-Peter<br />

et al. [14]. Amplified inserts were<br />

resolved in a 0.8% agarose gel, extracted using the<br />

Gene Clean II Kit, cloned in the vector pCR2.1, and<br />

subsequently transformed in E. coli InvF cells<br />

(Invitrogen) for DNA sequencing. The FLbR<br />

cDNA inserts were fully-sequenced in both directions<br />

at the DNA Sequencing Facility <strong>of</strong> the University<br />

<strong>of</strong> Nebraska-Lincoln. Cowpea FLbR<br />

nucleotide sequence and the deduced polypeptide<br />

sequence were searched for similarity in databases<br />

(GeneBank, EMBL, SwissProt) using programs <strong>of</strong><br />

the GCG package (Wisconsin Computer Group,<br />

version 8.0).<br />

3. Results and discussion<br />

3.1. Purification <strong>of</strong> <strong>cowpea</strong> FLbR<br />

FLbR was purified to homogeneity <strong>from</strong> crude<br />

extracts <strong>of</strong> <strong>cowpea</strong> root nodules by a four-step<br />

procedure involving ammonium sulfate precipitation,<br />

hydroxylapatite, ion exchange and size-exclusion<br />

chromatography. The purified <strong>cowpea</strong> FLbR<br />

had a specific activity <strong>of</strong> 216 nmol Lb 2+ O 2 min −1<br />

mg −1 <strong>of</strong> protein, which corresponded to a purification<br />

<strong>of</strong> approximately 1000-fold and a yield <strong>of</strong> 16%<br />

(Table 1). The hydroxylapatite column was an<br />

important step in the purification which resulted a<br />

60-fold increase in specific activity <strong>of</strong> FLbR al-


164<br />

though approximately 50% <strong>of</strong> the total activity<br />

was lost. A 40-fold decrease in the ratio <strong>of</strong> diaphorase<br />

to FLbR activity (Table 1) indicated that<br />

many <strong>of</strong> the other contaminating diaphorases were<br />

removed at this step. On the FPLC anion exchange<br />

column, FLbR was eluted as a sharp peak<br />

at about 15% <strong>of</strong> NaCl gradient (150 mM NaCl,<br />

Fig. 1). The specific activity <strong>of</strong> the purified <strong>cowpea</strong><br />

FLbR was about 50% <strong>of</strong> that reported for the<br />

soybean enzyme [7]. The difference in the specific<br />

activities between the <strong>cowpea</strong> and soybean FLbRs<br />

could result <strong>from</strong> inherent differences in the two<br />

enzymes.<br />

P. Luan et al. / Plant Science 154 (2000) 161–170<br />

Table 1<br />

Purification and specific activities <strong>of</strong> FLbR <strong>from</strong> <strong>cowpea</strong> nodule cytosol a<br />

3.2. Homogeneity and molecular weight analysis<br />

<strong>of</strong> <strong>cowpea</strong> FLbR<br />

Samples collected <strong>from</strong> each purification step<br />

were subjected to 10% SDS-PAGE and silver<br />

staining (Fig. 2). The sample after the Sephacryl<br />

S-200 step (lane 4) exhibited a single distinct<br />

protein band <strong>of</strong> about 55 kDa. The molecular<br />

weight for native <strong>cowpea</strong> FLbR was estimated to<br />

be 110 kDa using the Sephacryl S-200 SF column<br />

(data not shown) and thus appears to be a homodimer.<br />

These molecular weights were similar to<br />

those reported for soybean FLbR [7].<br />

Steps Total protein DCIP <strong>reductase</strong> (A) specific activity<br />

Ratio A/Bd FLbR (B) specific activity Total FLbR<br />

(mg) (U mg−1 ) c<br />

(U mg−1 ) b activity (U)<br />

Crude extract 893 187 0.22 194<br />

850<br />

G-25 607 254 0.29 176 875<br />

Hydroxylapatite<br />

4.5 368<br />

17.3<br />

78<br />

21.3<br />

Mono-Q 0.31 2289 162 50 14.1<br />

S-200 0.15 2565<br />

216 32 11.9<br />

a Purification steps are described in Section 2.<br />

b One unit is defined as 1 nmol <strong>of</strong> DCPIP reduced per min.<br />

c One unit is defined as 1 nmol <strong>of</strong> Lb 2+ O2 formed per min.<br />

d This is the ratio <strong>of</strong> DCPIP <strong>reductase</strong> (diaphorase) specific activity to FLbR specific activity.<br />

Fig. 1. Separation <strong>of</strong> FLbR by an ion-exchange FPLC on a Mono-Q HR5/5 column. The column was equilibrated with Buffer<br />

A (50 mM Tris–HCl, pH 7.5), and eluted with a linear NaCl gradient <strong>from</strong> 0 to 35% <strong>of</strong> Buffer B (1 M NaCl, 50 mM Tris–HCl,<br />

pH 7.5) in 50 ml total volume (Buffer A plus Buffer B) at a flow rate <strong>of</strong> 1 ml min −1 . Cowpea FLbR was eluted at about 15%<br />

<strong>of</strong> the gradient (150 mM NaCl, as indicated by the arrow).


Fig. 2. SDS-Polyacrylamide gel electrophoresis <strong>of</strong> <strong>cowpea</strong><br />

FLbR fractions during purification. The gel was silver-stained<br />

to detect proteins. Lane 1, G-25 fraction (50 g protein); lane<br />

2, hydroxylapatite fraction (10 g protein); lane 3, Mono-Q<br />

fraction (5 g protein); lane 4, Sephacryl S-200 fraction (3 g<br />

protein); and lane 5, 10-kDa ladder (10 g protein).<br />

P. Luan et al. / Plant Science 154 (2000) 161–170 165<br />

3.3. Characterization <strong>of</strong> <strong>cowpea</strong> FLbR<br />

The first 20 amino acids on the N-terminus <strong>of</strong><br />

the <strong>cowpea</strong> FLbR were determined to be: A-S-G-<br />

S-D-E-N-D-V-V-V-I-G-G-G-P-G-G-Y-V. When<br />

this sequence was compared to sequences deposited<br />

in the GCG database, it was found to be<br />

100% identical with soybean FLbR, and 95.2%<br />

identical with pea DLDH. This indicates that the<br />

FLbRs and DLDHs are probably highly conserved<br />

in legumes.<br />

The purified <strong>cowpea</strong> FLbR reduced both<br />

<strong>cowpea</strong> Lb 3+ and soybean Lb 3+ at comparable<br />

rates in the presence <strong>of</strong> NADH, forming Lb 2+ O 2<br />

under aerobic conditions. The enzyme had maximum<br />

Lb 3+ reduction activity at pH 6.5, and had<br />

no activity at pH values below pH 4.8 or above<br />

pH 8.5. Reactions as a function <strong>of</strong> time were<br />

monitored spectrometrically for the reduction <strong>of</strong><br />

Fig. 3. Reduction <strong>of</strong> <strong>cowpea</strong> Lb 3+ by FLbR. Reaction mixture contained 50 mM potassium phosphate buffer, pH 6.5, 500 M<br />

NADH, 50 M <strong>cowpea</strong> Lb 3+ , and 2 g purified <strong>cowpea</strong> FLbR. The reaction was carried out in a 1-ml cuvette at room<br />

temperature, and the spectra were scanned at 5-min intervals.


166<br />

Table 2<br />

Kinetic properties <strong>of</strong> <strong>cowpea</strong> FLbR, soybean FLbR and pig<br />

DLDH<br />

P. Luan et al. / Plant Science 154 (2000) 161–170<br />

Vmax Kcat (s−1 )<br />

(U mg−1 ) (M−1s−1 )<br />

K m (M) K cat/K m<br />

Cowpea FLbR<br />

10.4 221a Cowpea<br />

Lb<br />

3.1 298<br />

3+<br />

Soybean 12.4 185a 2.5 201<br />

Lb 3+<br />

NADH 57 NA NA NA<br />

Soybean FLbR a,b<br />

Soybean 9.2 450 6.2 674<br />

Lb 3+<br />

NADH 46 16 000 220 4.8<br />

Lipoamide 716<br />

NA NA NA<br />

Pig DLDH b,c<br />

Soybean 28 350 1.1 40<br />

Lb 3+<br />

NADH 73 25 000 344 4.5<br />

Lipoamide 430<br />

NA NA NA<br />

a One enzyme unit is defined as 1 nmol <strong>of</strong> Lb 2+ O2 formed<br />

min−1 mg−1 .<br />

b Determined by Ji et al. [9].<br />

c One enzyme unit is defined as 1 nmol NADH oxidized<br />

min−1 mg−1 .<br />

<strong>cowpea</strong> Lb 3+ (Fig. 3) and soybean Lb 3+ . The<br />

absorption at 541 and 574 nm, which was contributed<br />

by Lb 2+ O 2, increased as a function <strong>of</strong><br />

time, whereas the absorption at 627 nm resulting<br />

<strong>from</strong> Lb 3+ decreased. Two isosbestic points at 525<br />

and 588 nm were present in the reduction <strong>of</strong><br />

<strong>cowpea</strong> Lb 3+ . The spectroscopic characteristics<br />

for the reduction <strong>of</strong> <strong>cowpea</strong> Lb 3+ by <strong>cowpea</strong><br />

FLbR were similar to those reported for the soybean<br />

enzyme [7,9].<br />

The K m and V max values <strong>of</strong> <strong>cowpea</strong> FLbR for<br />

<strong>cowpea</strong> Lb 3+ reduction were determined to be<br />

10.4 M and 221 U mg −1 , respectively. The corresponding<br />

K m and V max values for soybean Lb 3+<br />

reduction by <strong>cowpea</strong> FLbR were determined to be<br />

12.4 M and 185 U mg −1 , respectively. The K m<br />

value for NADH by the <strong>cowpea</strong> FLbR was determined<br />

to be 57 M (Table 2). These values are<br />

similar to those reported for soybean FLbR [7,9].<br />

The K cat (6.2 s −1 ) and K cat/K m values (674 M −1<br />

s −1 ) <strong>of</strong> soybean enzyme were about tw<strong>of</strong>old<br />

greater than the <strong>cowpea</strong> enzyme, 3.1 s −1 and 298<br />

M −1 s −1 , respectively (Table 2). The K m value <strong>of</strong><br />

pig DLDH for soybean Lb 3+ (28 M) was more<br />

than tw<strong>of</strong>old higher than those <strong>of</strong> <strong>cowpea</strong> FLbR<br />

(12.4 M) and soybean FLbR (9.2 M) for Lb 3+<br />

(Table 2). The catalytic efficiencies (K cat/K m) <strong>of</strong><br />

<strong>cowpea</strong> FLbR for <strong>cowpea</strong> Lb 3+ (298 M −1 s −1 )<br />

and soybean FLbR for soybean Lb 3+ (674 M −1<br />

s −1 ) were about eight- and 17-fold higher than<br />

that <strong>of</strong> pig DLDH (40 M −1 s −1 ) (Table 2). Conversely,<br />

the affinity <strong>of</strong> pig DLDH for lipoamide<br />

was higher than the two FLbRs. These data suggest<br />

that dehydrogenation <strong>of</strong> lipoamide is most<br />

efficiently catalyzed by DLDH, and reduction <strong>of</strong><br />

Lb 3+ is most efficiently catalyzed by FLbRs.<br />

Thus, although DLDH and FLbRs exhibit many<br />

similarities in their enzyme kinetics, they are expected<br />

to function differently in vivo.<br />

3.4. cDNA sequence <strong>of</strong> <strong>cowpea</strong> FLbR<br />

A cDNA fragment <strong>of</strong> 802 bp was amplified<br />

<strong>from</strong> a <strong>cowpea</strong> root nodule cDNA library by PCR<br />

using the FLbR Fwd and FLbR Rvs primers.<br />

DNA sequencing showed that the 802-bp PCRfragment<br />

encoded for a <strong>cowpea</strong> FLbR. Thus, this<br />

fragment was PCR-labelled by incorporating Dig-<br />

11-dUTP [15] and used as probe for screening the<br />

above <strong>cowpea</strong> cDNA library. About 5×10 5 recombinant<br />

phage plaques were screened, resulting<br />

in seven positive plaques that were numbered c1<br />

through c7. The inserts <strong>from</strong> these plaques were<br />

amplified by PCR using primers [14]. Inserts<br />

<strong>from</strong> plaques c4 and c6 (named clones 4 and 6)<br />

were about 1.8 kb in length and thus they were<br />

subcloned for DNA sequencing. Sequence comparison<br />

showed that clones 4 and 6 are copies <strong>of</strong><br />

the same cDNA, and that they code for the same<br />

protein. Comparison with sequences deposited in<br />

the GenBank database revealed that clones 4 and<br />

6 have high similarity (80%, see below) to the<br />

soybean FLbR, and other DHLD sequences, and<br />

thus that they code for a <strong>cowpea</strong> FLbR.<br />

The complete <strong>cowpea</strong> FLbR cDNA consists <strong>of</strong><br />

1797 nucleotides with an open reading frame <strong>of</strong><br />

1569 bases (Fig. 4). The poly(A + )11 tail is present<br />

150 bases after the stop codon position. The deduced<br />

polypeptide sequence has 523 amino acids<br />

with a predicted molecular weight <strong>of</strong> 56 kDa,<br />

which corresponds well to the observed value <strong>of</strong> 55


P. Luan et al. / Plant Science 154 (2000) 161–170 167<br />

Fig. 4. Nucleotide and deduced amino acid sequences <strong>of</strong> <strong>cowpea</strong> FLbR. The experimentally determined N-terminal amino acid<br />

sequence <strong>of</strong> the purified enzyme is underlined in the amino acid sequence. The deduced leader sequence is indicated in italics<br />

(amino acid residues 1–30). The cysteine residues hypothesized to form the active disulfide bond are asterisked.


Table 3<br />

Conserved domains in <strong>cowpea</strong> FLbR a<br />

Conserved sequences in the FAD-binding domain<br />

FLbR-<strong>cowpea</strong> 37–66 N D V V V I G G G P G G Y V A A I K A S Q L G L K T T C I E<br />

FLbR-soybean · · · · · · · · · · · · · · · · A<br />

b 37–66 · · ·<br />

· · · · · · · · · ·<br />

DLDH-pea I · · · · ·<br />

b 38–67 · · · ·<br />

· · · · · · · · · A · · · F · · · · · ·<br />

DLDH-human 42–71 T · · · S · · · · · · · · · · · A · · · F · · V · · ·<br />

b A · ·<br />

DLDH-yeast 28–56 – · · · I · · · · · A · · · · · · · · A · · · F N · A · V ·<br />

b<br />

DLDH-E. coli · L · · · · A<br />

b 6–35 T Q · ·<br />

· · S · · F R C A D · · · E · V I V ·<br />

GSHR-human L · · · · · S · · L A S · R R · A<br />

b 21–50 Y · Y<br />

E · · A R A A V V ·<br />

GSHR-E. coli b 5–34 T · T I A · · · · S · · I A S I N R · A M Y · Q · C A L · ·<br />

Conserved sequences in the disulfide active site<br />

L G G T CcL N V G Cc FLbR-<strong>cowpea</strong> 67–96 K R G T<br />

I P S K A L L H S S H M Y H E A<br />

FLbR-soybean 67–96 · · · · · · · · · · · · · · · · · · · · · · · · · · · · · ·<br />

DLDH-pea 69–98 · · · A · · · · · · · · · · · · · · · · · · · · · · · · · ·<br />

DLDH-human 73–102 · N E · · · · · · · · · · · · · · · · · · N N · · Y · · M ·<br />

DLDH-yeast 58–87 · · · K · · · · · · · · · · · · · · · · · N N · · L F · Q M<br />

DLDH-E. coli 37–66 R Y N · · · · V · · · · · · · · · · · · · · V A K V I E · ·<br />

GSHR-human 52–80 S H K – · · · · · V · · · · V · · · T M W N T A V H S E F M<br />

GSHR-E. coli 36–64 · N E – · · · · · V · · · · V · K · V M W · A A Q I R E A I<br />

Conserved sequences in the NAD(P)H domain<br />

FLbR-<strong>cowpea</strong> 196–225 S S T G A L A L T E I P K K L V V I G A G Y I G L E M G S V<br />

FLbR-soybean 196–225 · · · · · · · · S · · · · · · · · · · · · · · · · · · · · ·<br />

DLDH-pea 198–227 · · · · · · · · S · · · · · · · · · · · · · · · · · · · · ·<br />

DLDH-human 203–232 · · · · · · S · K K V · E · M · · · · · · V · · V · L · · ·<br />

DLDH-yeast 194–223 · · · · · · S · K · · · · R · T I · · G · I · · · · · · · ·<br />

DLDH-E. coli 163–192 D · · S · · E · K · V · E R · L · M · G · I · · · · · · T ·<br />

GSHR-human 173–202 D · D · F F · · P A L · E R V A · V · · · · · A V · L A G ·<br />

GSHR-E. coli 161–190 T · D · F F Q · E · L · G R S · I V · S · · · A V · · A G I<br />

a Identical amino acids are shown as dots, gaps are shown as dashes.<br />

b Sequences are cited <strong>from</strong> Ji et al. [8].<br />

c Hypothesized disulfide cysteines.<br />

168<br />

P. Luan et al. / Plant Science 154 (2000) 161–170


kDa. By comparing the deduced N-terminal sequence<br />

to the mature protein sequence, we found<br />

the existence <strong>of</strong> a 32-amino acid leader peptide.<br />

This leader peptide is rich in basic and hydroxylated<br />

amino acids, and deficient in acidic residues<br />

which is highly similar to that <strong>of</strong> soybean FLbR<br />

[9]: the first 20 amino acids are identical, and only<br />

four amino acids are different in the last 10 amino<br />

acids.<br />

Comparison <strong>of</strong> the sequence <strong>of</strong> <strong>cowpea</strong> FLbR<br />

cDNA to known sequences revealed striking similarities<br />

to soybean FLbR cDNA [8] (88%), pea<br />

DLDH (EC 1.8.1.4) gene [16] (85%) and other<br />

DLDH genes [17,18], glutathione <strong>reductase</strong> (EC<br />

1.6.4.2) [19], and mercuric <strong>reductase</strong> (EC 1.16.1.1)<br />

[20] (20–60%). All <strong>of</strong> these enzymes belong to the<br />

pyridine nucleotide-disulfide oxido-<strong>reductase</strong> family<br />

[21]. Of the 523 deduced amino acids <strong>of</strong> <strong>cowpea</strong><br />

FLbR, 445 were identical to soybean FLbR [8];<br />

441 identical to pea DLDH [16]; 275 identical or<br />

less to other enzymes in this family [17–20].<br />

Pileup (GCG package) analyses <strong>of</strong> the amino<br />

acid sequences for the FLbR and the pyridine<br />

nucleotide-disulfide oxido-<strong>reductase</strong> family enzymes<br />

[21]showed that important residues and<br />

functional domains for FAD-binding, disulfide active<br />

site and NAD(P)H-binding were highly conserved<br />

(Table 3). The FAD-binding domain in<br />

<strong>cowpea</strong> FLbR is essentially identical (residue 37–<br />

66) to soybean FLbR, only three residues different<br />

<strong>from</strong> pea DLDH, and 6–18 residues different<br />

<strong>from</strong> the other enzymes in this class (Table 3). The<br />

disulfide-active site <strong>of</strong> <strong>cowpea</strong> FLbR is proposed<br />

to be located <strong>from</strong> residue 67 to 96, which is<br />

identical to soybean FLbR, one amino acid different<br />

<strong>from</strong> pea DLDH, and 6–17 amino acids different<br />

<strong>from</strong> the others. The region for the<br />

NAD(P)H-binding domain in <strong>cowpea</strong> FLbR is<br />

identified <strong>from</strong> residue 196 to 225, which is one<br />

residue different <strong>from</strong> soybean FLbR and pea<br />

DLDH, and 9–18 residues different <strong>from</strong> the others.<br />

All <strong>of</strong> these sites are located on the N-terminal<br />

half <strong>of</strong> the protein. The presence <strong>of</strong> a high homology<br />

in the functional domains and c<strong>of</strong>actor-binding<br />

domains suggests a similar origin and enzyme<br />

mechanism for these proteins. However, the kinetic<br />

constants and catalytic efficiencies for Lb 3+<br />

reduction by the FLbRs are very different <strong>from</strong><br />

DLDHs, indicating that the two are not identical<br />

and may have different mechanisms. The existence<br />

<strong>of</strong> FLbR in <strong>cowpea</strong> and soybean suggests that this<br />

P. Luan et al. / Plant Science 154 (2000) 161–170 169<br />

enzyme may be common to all legumes. As hypothesized<br />

for soybean FLbR [6,11], FLbRs in<br />

<strong>cowpea</strong> and other legumes probably reduce Lb 3+<br />

in vivo to maintain adequate levels <strong>of</strong> functional<br />

Lb 2+ form.<br />

Acknowledgements<br />

This work was supported in part by Grants<br />

<strong>from</strong> the National Science Foundation (no. OSR-<br />

92552255), and the U.S. Department <strong>of</strong> Agriculture<br />

Cooperative State Research, Education and<br />

Extension Service (no. 95-37305-2441) to R.V.<br />

Klucas, the Center for Biotechnology, University<br />

<strong>of</strong> Nebraska-Lincoln funded through the Nebraska<br />

Research Initiative to Gautam Sarath, and<br />

the Consejo Nacional de Ciencia y Tecnología<br />

(project number 25229-N), México, to Raúl<br />

Arredondo-Peter.<br />

References<br />

[1] C.A. Appleby, Leghemoglobin and Rhizobium respiration,<br />

Annu. Rev. Plant Physiol. 35 (1984) 443–478.<br />

[2] M. Becana, R.V. Klucas, Oxidation and reduction <strong>of</strong><br />

<strong>leghemoglobin</strong> in root nodules <strong>of</strong> leguminous plants,<br />

Plant Physiol. 98 (1992) 1217–1221.<br />

[3] C.A. Appleby, Properties <strong>of</strong> leghaemoglobin in io, and<br />

its isolation as ferrous oxyleghaemoglobin, Biochim. Biophys.<br />

Acta 188 (1969) 222–229.<br />

[4] V.L. Kretovich, S.S. Melik-Sarkisyan, N.F. Bashirova,<br />

A.F. Topunov, Enzymatic reduction <strong>of</strong> <strong>leghemoglobin</strong> in<br />

lupin nodules, J. Appl. Biochem. 4 (1982) 209–217.<br />

[5] L.I. Golubeva, A.F. Topunov, S.S. Goncharov, K.B.<br />

Aseeva, V.L. Kretovich, Production and properties <strong>of</strong> a<br />

homogeneous preparation <strong>of</strong> metleghemogloin <strong>reductase</strong><br />

<strong>of</strong> lupine root nodule cytosol, Biokhimiya 53 (1998)<br />

1478–1482 English translation.<br />

[6] L.L. Saari, R.V. Klucas, Ferric <strong>leghemoglobin</strong> <strong>reductase</strong><br />

<strong>from</strong> soybean root nodules, Arch. Biochem. Biophys. 231<br />

(1984) 102–113.<br />

[7] L. Ji, S. Wood, M. Becana, R.V. Klucas, Purification<br />

and characterization <strong>of</strong> soybean root nodule <strong>ferric</strong><br />

<strong>leghemoglobin</strong> <strong>reductase</strong>, Plant Physiol. 96 (1994) 32–37.<br />

[8] L. Ji, M. Becana, G. Sarath, R.V. Klucas, Cloning and<br />

sequence analysis <strong>of</strong> a cDNA encoding <strong>ferric</strong><br />

<strong>leghemoglobin</strong> <strong>reductase</strong> <strong>from</strong> soybean nodules, Plant<br />

Physiol. 104 (1994) 453–459.<br />

[9] L. Ji, M. Becana, G. Sarath, L. Shearman, R.V. Klucas,<br />

Overproduction in Escherichia coli and characterization<br />

<strong>of</strong> soybean <strong>ferric</strong> <strong>leghemoglobin</strong> <strong>reductase</strong>, Plant Physiol.<br />

106 (1994) 203–209.<br />

[10] M. Becana, M.L. Salin, L. Ji, R.V. Klucas, Flavin-mediated<br />

reduction <strong>of</strong> <strong>ferric</strong> <strong>leghemoglobin</strong> <strong>from</strong> soybean<br />

nodules, Planta 183 (1991) 575–583.


170<br />

[11] H.K. Jun, G. Sarath, F.W. Wagner, Detection and<br />

purification <strong>of</strong> modified <strong>leghemoglobin</strong>s <strong>from</strong> soybean<br />

root nodules, Plant Sci. 100 (1994) 31–40.<br />

[12] M.G. Murray, C. Ross, Molecular weight estimations <strong>of</strong><br />

some pyrimidine-metabolizing enzymes <strong>from</strong> pea cotyledons<br />

by gel filtration, Phytochemistry 10 (1971) 2645–<br />

2648.<br />

[13] J. Sambrook, E.F. Fritsch, T. Maniatis, Molecular<br />

Cloning: A Laboratory Manual, Cold Spring Harbor<br />

Laboratory, Cold Spring Harbor, NY, 1989, pp. 2.60–<br />

2.121.<br />

[14] R. Arredondo-Peter, A. Bonic, G. Sarath, R.V. Klucas,<br />

Rapid PCR-based detection <strong>of</strong> inserts <strong>from</strong> cDNA libraries<br />

using phage pools or direct phage plaques and<br />

lambda primers, Plant Mol. Biol. Rep. 13 (1991) 138–<br />

146.<br />

[15] Y.H. Lu, S. Negré, P. Leroy, M. Bernard, PCR-mediated<br />

screening and labeling <strong>of</strong> DNA <strong>from</strong> clones, Plant Mol.<br />

Biol. Rep. 11 (1993) 345–349.<br />

[16] S.R. Turner, R. Ireland, S. Rawsthorne, Purification and<br />

primary amino acid sequence <strong>of</strong> the L subunit <strong>of</strong> glycine<br />

decarboxylase. Evidence for a single lipoamide dehydrogenase<br />

in plant mitochondria, J. Biol. Chem. 267 (1992)<br />

7745–7750.<br />

P. Luan et al. / Plant Science 154 (2000) 161–170<br />

.<br />

[17] P.E. Stephen, H.M. Lewis, M.G. Darlison, J.R. Guest,<br />

Nucleotide sequence <strong>of</strong> the lipoamide dehydrogenase <strong>of</strong><br />

E. coli K12, Eur. J. Biochem. 135 (1983) 519–527.<br />

[18] J. Bourguignon, D. Macherel, M. Neuburger, R. Douce,<br />

Isolation, characterization, and sequence analysis <strong>of</strong> a<br />

cDNA clone encoding L-protein, the dihydrolipoamide<br />

dehydrogenase component <strong>of</strong> the glycine cleavage system<br />

<strong>from</strong> pea-leaf mitochondria, Eur. J. Biochem. 204 (1992)<br />

865–873.<br />

[19] R.L. Krauth-Siegel, R. Blatterspiel, M. Saleh, E. Schilts,<br />

R.H. Schirmer, R. Untucht-Grau, Glutathione <strong>reductase</strong><br />

<strong>from</strong> human erythrocytes, the sequence <strong>of</strong> the NADPH<br />

domain and <strong>of</strong> the interface domain, Eur. J. Biochem.<br />

121 (1982) 259–267.<br />

[20] R.A. Laddaga, L. Chu, T.K. Misra, S. Silver, Nucleotide<br />

sequence and expression <strong>of</strong> the mercuric-resistance<br />

operon <strong>from</strong> Staphylococcus aureus plasmid pI258, Proc.<br />

Natl. Acad. Sci. USA 84 (1987) 5106–5110.<br />

[21] C.H. Williams, Lipoamide dehydrogenase, glutathione<br />

<strong>reductase</strong>, thioredoxin <strong>reductase</strong> and mercuric ion <strong>reductase</strong>-family<br />

<strong>of</strong> flavoenzyme transhydrogenases, in: F.<br />

Muller (Ed.), Chemistry and Biochemistry <strong>of</strong> Flavoenzymes,<br />

CRC Press, Boca Raton, FL, 1992, pp. 121–211.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!