8
JOURNAL OF BACTERIOLOGY, 0021-9193/00/$04.0010 Oct. 2000, p. 5440–5447 Vol. 182, No. 19 Copyright © 2000, American Society for Microbiology. All Rights Reserved. The orf162b Sequence of Rhodobacter capsulatus Encodes a Protein Required for Optimal Levels of Photosynthetic Pigment-Protein Complexes MUKTAK AKLUJKAR, 1 ANDREA L. HARMER, 1 ROGER C. PRINCE, 2 AND J. THOMAS BEATTY 1 * Department of Microbiology and Immunology, University of British Columbia, Vancouver, British Columbia, Canada, 1 and ExxonMobil Research and Engineering Company, Annandale, New Jersey 2 Received 26 April 2000/Accepted 10 July 2000 The orf162b sequence, the second open reading frame 3* of the reaction center (RC) H protein gene puhA in the Rhodobacter capsulatus photosynthesis gene cluster, is shown to be transcribed from a promoter located 5* of puhA. A nonpolar mutation of orf162b was generated by replacing most of the coding region with an antibiotic resistance cartridge. Although the mutant strain initiated rapid photosynthetic growth, growth slowed progressively and cultures often entered a pseudostationary phase. The amounts of the RC and light harvesting complex I (LHI) in cells obtained from such photosynthetic cultures were abnormally low, but these deficiencies were less severe when the mutant was grown to a pseudostationary phase induced by low aeration in the absence of illumination. The orf162b mutation did not significantly affect the expression of a pufB::lacZ translationally in-frame gene fusion under the control of the puf promoter, indicating normal transcription and translation of RC and LHI genes. Spontaneous secondary mutations in the strain with the orf162b disruption resulted in a bypass of the photosynthetic growth retardation and reduced the level of light harvesting complex II. These results and the presence of sequences similar to orf162b in other species indicate that the Orf162b protein is required for normal levels of the photosynthetic apparatus in purple photosynthetic bacteria. Purple nonsulfur photosynthetic bacteria such as Rho- dobacter capsulatus are capable of aerobic respiratory and an- aerobic photosynthetic growth. The photosynthetic apparatus includes three membrane-bound pigment-protein complexes: the reaction center (RC), where light-dependent electron transfer is initiated; light harvesting (LH) complex I, which is adjacent to and perhaps forms a ring encircling the RC as part of the so-called core complex; and the LHII complex, which is thought to be present in multiple copies of a ring-shaped struc- ture that interconnect core complexes (20). These complexes are located within differentiated invaginations of the cytoplas- mic membrane called the intracytoplasmic membrane system (ICM), which is formed upon oxygen deprivation of cultures (12). The presence of the various photosynthetic complexes can be evaluated by their characteristic light absorption spec- tra, which reflect the protein environments around bacterio- chlorophyll a (Bchl). For example, the Bchl’s of the LHII complex of R. capsulatus absorb light of 800 and 850 nm, whereas the Bchl’s of the less abundant LHI complex absorb approximately 870-nm light (13). Two of the three protein subunits of the RC, designated RC L and RC M, and both protein subunits of LHI (LHI a and LHI b), are encoded by the puf operon (2). The third subunit of the RC, called RC H, is encoded by the puhA gene, which is transcribed as part of the bchFNBHLM-lhaA-puhA super- operon from two promoters, one 59 of bchF and the other within the lhaA gene (3). As shown in Fig. 1, several open reading frames (ORFs) located 39 of puhA and in the same transcriptional orientation have been identified on the basis of DNA sequence analysis (2). A previous publication reported that disruptions of orf214, located immediately 39 of the puhA gene, resulted in reduced amounts of the RC and LHI com- plexes and abolished photosynthetic growth, and it was sug- gested that the Orf214 protein is an RC assembly factor (28). The orf162b sequence immediately follows orf214 as the second ORF 39 of puhA. ORFs similar to orf162b have been found in four other species of purple photosynthetic bacteria: orf153 in Rhodobacter sphaeroides (M. Choudhary and S. Kaplan, personal communication), orf154 in Rubrivivax gelati- nosus (K. Nagashima, personal communication), orfI3087 in Rhodospirillum rubrum (6), and orf168 in Rhodopseudomonas palustris (genome sequence made available by the Joint Ge- nome Institute at http://spider.jgi-psf.org/JGI_microbial/html /rhodo_homepage.html). In all five species the ORF similar to orf162b is immediately 39 of a homologue of orf214, which immediately follows the puhA gene, just as in R. capsulatus (2, 6, 9a; K. Nagashima, personal communication; M. Aklujkar, analysis of the R. palustris genome). The predicted protein sequences are only 43% (R. sphaeroides), 15% (R. gelatinosus), 14% (R. rubrum), and 17% (R. palustris) identical to Orf162b, but they have similar hydropathy profiles (using the Goldman- Engelman-Steitz algorithm and the TOPPRED program [10]) with a transmembrane segment near the amino terminus. None of these predicted proteins has significant sequence sim- ilarity to proteins of known function. In this paper we present evidence that orf162b encodes a protein that is required for optimal photosynthetic growth, and that a disruption of orf162b reduces the amounts of RC and LHI in oxygen-deprived and photosynthetically grown cells. This phenotype is complemented in trans by a plasmid-borne copy of orf162b. We also demonstrate that transcription of orf162b is abolished by insertion of a transcription termination sequence into the puhA gene, and that the orf162b mutant phenotype is suppressed by spontaneous mutations that reduce the amount of LHII. * Corresponding author. Mailing address: Department of Microbi- ology and Immunology, University of British Columbia, 300 - 6174 University Blvd., Vancouver, British Columbia, Canada V6T 1Z3. Phone: (604) 822-6896. Fax: (604) 822-6041. E-mail: jbeatty@interchange .ubc.ca. 5440 on April 5, 2021 by guest http://jb.asm.org/ Downloaded from

orf162b Sequence of Encodes a Protein Required for Optimal ...MUKTAK AKLUJKAR,1 ANDREA L. HARMER,1 ROGER C. PRINCE,2 AND J. THOMAS BEATTY1* Department of Microbiology and Immunology,

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

  • JOURNAL OF BACTERIOLOGY,0021-9193/00/$04.0010

    Oct. 2000, p. 5440–5447 Vol. 182, No. 19

    Copyright © 2000, American Society for Microbiology. All Rights Reserved.

    The orf162b Sequence of Rhodobacter capsulatus Encodes aProtein Required for Optimal Levels of Photosynthetic

    Pigment-Protein ComplexesMUKTAK AKLUJKAR,1 ANDREA L. HARMER,1 ROGER C. PRINCE,2 AND J. THOMAS BEATTY1*

    Department of Microbiology and Immunology, University of British Columbia, Vancouver, British Columbia, Canada,1

    and ExxonMobil Research and Engineering Company, Annandale, New Jersey2

    Received 26 April 2000/Accepted 10 July 2000

    The orf162b sequence, the second open reading frame 3* of the reaction center (RC) H protein gene puhA inthe Rhodobacter capsulatus photosynthesis gene cluster, is shown to be transcribed from a promoter located 5*of puhA. A nonpolar mutation of orf162b was generated by replacing most of the coding region with anantibiotic resistance cartridge. Although the mutant strain initiated rapid photosynthetic growth, growthslowed progressively and cultures often entered a pseudostationary phase. The amounts of the RC and lightharvesting complex I (LHI) in cells obtained from such photosynthetic cultures were abnormally low, but thesedeficiencies were less severe when the mutant was grown to a pseudostationary phase induced by low aerationin the absence of illumination. The orf162b mutation did not significantly affect the expression of a pufB::lacZtranslationally in-frame gene fusion under the control of the puf promoter, indicating normal transcription andtranslation of RC and LHI genes. Spontaneous secondary mutations in the strain with the orf162b disruptionresulted in a bypass of the photosynthetic growth retardation and reduced the level of light harvesting complexII. These results and the presence of sequences similar to orf162b in other species indicate that the Orf162bprotein is required for normal levels of the photosynthetic apparatus in purple photosynthetic bacteria.

    Purple nonsulfur photosynthetic bacteria such as Rho-dobacter capsulatus are capable of aerobic respiratory and an-aerobic photosynthetic growth. The photosynthetic apparatusincludes three membrane-bound pigment-protein complexes:the reaction center (RC), where light-dependent electrontransfer is initiated; light harvesting (LH) complex I, which isadjacent to and perhaps forms a ring encircling the RC as partof the so-called core complex; and the LHII complex, which isthought to be present in multiple copies of a ring-shaped struc-ture that interconnect core complexes (20). These complexesare located within differentiated invaginations of the cytoplas-mic membrane called the intracytoplasmic membrane system(ICM), which is formed upon oxygen deprivation of cultures(12). The presence of the various photosynthetic complexescan be evaluated by their characteristic light absorption spec-tra, which reflect the protein environments around bacterio-chlorophyll a (Bchl). For example, the Bchl’s of the LHIIcomplex of R. capsulatus absorb light of 800 and 850 nm,whereas the Bchl’s of the less abundant LHI complex absorbapproximately 870-nm light (13).

    Two of the three protein subunits of the RC, designated RCL and RC M, and both protein subunits of LHI (LHI a andLHI b), are encoded by the puf operon (2). The third subunitof the RC, called RC H, is encoded by the puhA gene, which istranscribed as part of the bchFNBHLM-lhaA-puhA super-operon from two promoters, one 59 of bchF and the otherwithin the lhaA gene (3). As shown in Fig. 1, several openreading frames (ORFs) located 39 of puhA and in the sametranscriptional orientation have been identified on the basis ofDNA sequence analysis (2). A previous publication reported

    that disruptions of orf214, located immediately 39 of the puhAgene, resulted in reduced amounts of the RC and LHI com-plexes and abolished photosynthetic growth, and it was sug-gested that the Orf214 protein is an RC assembly factor (28).

    The orf162b sequence immediately follows orf214 as thesecond ORF 39 of puhA. ORFs similar to orf162b have beenfound in four other species of purple photosynthetic bacteria:orf153 in Rhodobacter sphaeroides (M. Choudhary and S.Kaplan, personal communication), orf154 in Rubrivivax gelati-nosus (K. Nagashima, personal communication), orfI3087 inRhodospirillum rubrum (6), and orf168 in Rhodopseudomonaspalustris (genome sequence made available by the Joint Ge-nome Institute at http://spider.jgi-psf.org/JGI_microbial/html/rhodo_homepage.html). In all five species the ORF similar toorf162b is immediately 39 of a homologue of orf214, whichimmediately follows the puhA gene, just as in R. capsulatus (2,6, 9a; K. Nagashima, personal communication; M. Aklujkar,analysis of the R. palustris genome). The predicted proteinsequences are only 43% (R. sphaeroides), 15% (R. gelatinosus),14% (R. rubrum), and 17% (R. palustris) identical to Orf162b,but they have similar hydropathy profiles (using the Goldman-Engelman-Steitz algorithm and the TOPPRED program [10])with a transmembrane segment near the amino terminus.None of these predicted proteins has significant sequence sim-ilarity to proteins of known function.

    In this paper we present evidence that orf162b encodes aprotein that is required for optimal photosynthetic growth, andthat a disruption of orf162b reduces the amounts of RC andLHI in oxygen-deprived and photosynthetically grown cells.This phenotype is complemented in trans by a plasmid-bornecopy of orf162b. We also demonstrate that transcription oforf162b is abolished by insertion of a transcription terminationsequence into the puhA gene, and that the orf162b mutantphenotype is suppressed by spontaneous mutations that reducethe amount of LHII.

    * Corresponding author. Mailing address: Department of Microbi-ology and Immunology, University of British Columbia, 300 - 6174University Blvd., Vancouver, British Columbia, Canada V6T 1Z3.Phone: (604) 822-6896. Fax: (604) 822-6041. E-mail: [email protected].

    5440

    on April 5, 2021 by guest

    http://jb.asm.org/

    Dow

    nloaded from

    http://jb.asm.org/

  • MATERIALS AND METHODS

    Bacterial strains and plasmids. The photosynthetically wild-type R. capsulatusstrain SB1003 (25), the gene transfer agent (GTA) overproducer strain DE442(29), and the puhA polar mutant DW1 (28) have been described previously.Escherichia coli C600 r2 m1 (7) and DH5a (Life Technologies, GIBCO BRL)were used for the construction and maintenance of plasmids; strains HB101(pRK2013) and TEC5 (11, 26) were used in E. coli-to-R. capsulatus conjugations.Plasmids pXCA601::935, pUC13, and pRR5C have been described previously (1,19, 30). Plasmid pUC4KIXX was from Pharmacia Biotech, Inc. Bacterial strainsand plasmids produced in this study are described below.

    Growth conditions and media. E. coli strains were grown in Luria-Bertani(LB) medium (22) or on LB agar plates. Highly aerated and oxygen-deprivedcultures of R. capsulatus were grown at 30°C without illumination in RCVmedium (5), in Erlenmeyer flasks filled to 20 or 80% of nominal capacity,respectively, and shaken at 300 or 150 rpm, respectively. Highly aerated culturesdid not enter the stationary phase until reaching about 300 Klett units (seebelow), whereas the growth of oxygen-deprived cultures began to slow at about80 Klett units. Photosynthetic cultures were grown anaerobically in screw-captubes (20 ml) or Roux bottles (800 ml) inoculated from oxygen-deprived culturesand filled with RCV, or on RCV agar plates placed in BBL GasPak anaerobicjars (Becton Dickinson & Co.). Photosynthetic cultures were incubated at 30°Cin an aquarium filled with water and illuminated by tungsten filament incandes-cent lamps at a light intensity of 150 microeinsteins/m2/s, measured with aphotometer equipped with the LI-190SB quantum sensor (LI-COR Inc.). Cul-ture turbidity was monitored with a Klett-Summerson photometer equipped witha red (no. 66) filter (100 Klett units 5 3.3 3 108 CFU/ml). Antibiotic-resistant E.coli and R. capsulatus strains were selected with 25 and 10 mg of kanamycin/ml,10 and 0.5 mg of tetracycline/ml, and 10 and 2 mg of gentamicin/ml, respectively.

    DNA techniques. Recombinant DNA procedures were carried out essentiallyas described previously (22). Plasmid DNA was isolated from cells and fromagarose gels with kits from QIAGEN. Conjugative transfer of plasmids from E.coli to R. capsulatus was performed by mixing of overnight cultures of donor,helper, and recipient cells in a 1:1:1 ratio by volume; centrifugation at 15,000 3g for 2 min; resuspension in an equal volume of RCV; and incubation of a 10-mlaliquot adsorbed onto an RCV agar plate overnight at 30°C. Donor cells wereabsent from the negative controls. Cells from each spot were streaked onto RCVagar plates containing the appropriate antibiotics. Transconjugant colonies werestreaked onto YPS agar (22) plates to ensure the absence of E. coli donors.

    Construction of the nonpolar orf162b mutant SBK1. A 2-kbp DNA fragmentfrom BamHI digestion of the R9 plasmid pBLM2 (26) containing the orf162bgene and flanking sequences (Fig. 1) was inserted into the BamHI site of pUC13,and 63% of the coding sequence of orf162b was eliminated by digestion withBsaBI and PflMI. T4 DNA polymerase was used to resect single-stranded pro-truding 39 ends from this deletion, and a NotI site in the linker (59 AGCGGCCGCT 39) was inserted by linker tailing (22); the resultant plasmid was cut at thisNotI site, and the ends were filled in with the Klenow fragment. A 1.3-kbp SmaIfragment containing the kanamycin resistance gene and the first 153 bp of thebleomycin resistance gene from pUC4KIXX was ligated into the filled-in NotIsite such that the transcriptional orientation of these genes is the same as inorf162b. This fragment does not usually terminate transcription (8), and itsnonpolar effect was confirmed by genetic complementation (see Results). E. coliTEC5 was transformed with the plasmid carrying the deleted, kanamycin resis-tance-marked orf162b gene. Homologous recombination of this pUC-derivedplasmid with the conjugative plasmid pDPT51 in E. coli TEC5 permitted con-

    jugative transfer of kanamycin resistance from TEC5 to R. capsulatus strainDE442. DE442 overproduces phage-like GTA particles containing random 4.6-kbp linear fragments of genomic DNA (29), which were used to replace thewild-type orf162b gene in SB1003 by transduction, producing strain SBK1.

    Construction of the complementation plasmid pAH8. A pUC13 constructcontaining the ClaI-to-EcoRI segment including orf162b (Fig. 1) was digestedwith PstI, and an EcoRI site in the linker (59 CCGAATTCGG 39) was ligatedinto this site (made blunt with T4 DNA polymerase) by linker tailing. The 748-bpEcoRI fragment containing orf162b was excised from the resultant plasmid andligated into the vector pRR5C (30). The resultant plasmid, called pAH8, whichcontains the orf162b gene transcribed from the puf promoter, was used to restoreexpression of orf162b in mutant backgrounds.

    RNA blots. Aerobic and oxygen-limited cultures were grown to a density of 100Klett units, and RNA was isolated from 25 ml of each culture using the RNeasykit (QIAGEN). Samples were treated with 30 U of DNase I in 100 mM sodiumacetate–5 mM MgSO4 (pH 5.0) for 30 min at room temperature, followed byphenol-chloroform extraction and ethanol precipitation. Seven micrograms ofeach RNA sample was used for formaldehyde gel electrophoresis (16) andelectroblotted onto a nylon membrane (ICN) for 2 h at 80 V in 0.53 Tris-borate-EDTA (TBE). The membrane was baked at 80°C for 2 h, placed on a nylonmesh, and prehybridized in 10 ml of 50% formamide–10% dextran sulfate–5.8%NaCl–1% sodium dodecyl sulfate (SDS)–0.2% bovine serum albumin (BSA)–0.2% Ficoll–0.2% polyvinylpyrrolidone–0.1 sodium pyrophosphate–50 mM TrisHCl (pH 7.5) with 0.1 mg of sheared salmon sperm DNA (Sigma)/ml at 42°C for3 h with rotation (in an oven from BIO/CAN Scientific). The probe was agel-purified DNA fragment extending from the BsaBI site in orf162b to theEcoRI site in orf55 (Fig. 1), labeled with 32P by using the Redi-Prime kit(Amersham Pharmacia Biotech) for 2 h. After 16 h of hybridization, the mem-brane was washed with agitation twice with 100 ml of 23 SSC (13 SSC is 0.15 MNaCl plus 0.015 M sodium citrate) for 10 min at room temperature, twice with100 ml of 23 SSC containing 1% SDS for 15 min at 60°C, and once with 100 mlof 0.13 SSC for 15 min at room temperature. Hybridization signals were de-tected with film (Kodak).

    b-Galactosidase assays. Cells from oxygen-deprived R. capsulatus cultures at80 or 165 Klett units, and photosynthetic cultures at 125 Klett units, wereharvested by centrifugation. Pellets from 40 ml of culture were stored at 280°Cbefore assay as previously described (18).

    Chromatophore isolation. Oxygen-deprived and photosynthetic culturesgrown to 150 Klett units were centrifuged, and the cell pellets were stored at280°C. The cells were resuspended in 20 ml of chromatophore buffer [20 mM3-(N-morpholino)propanesulfonate (MOPS), 100 mM KCl, 1 mM MgCl2 (pH7.2)] and passed through a French press three times. Cell debris was removed bycentrifugation at 12,000 3 g for 15 min, and the supernatant was centrifuged at412,000 3 g for 14 min to pellet membrane vesicles (chromatophores), whichwere resuspended in chromatophore buffer. An aliquot stored at 4°C was usedfor flash spectroscopy, and the remainder was stored at 280°C and used forprotein electrophoresis.

    Spectroscopy. Absorption spectroscopy of intact cells was performed as de-scribed previously (17), and data were collected with the Spectra Calc softwarepackage (Galactic Industries Corporation). Light scattering at 650 nm was usedto normalize the spectra. Low-temperature absorption spectroscopy used a Hi-tachi 557 double-beam spectrophotometer. Chromatophores in chromatophorebuffer (see above) were mixed with an equal volume of anhydrous glycerol andfrozen in liquid nitrogen. Spectra were obtained with the samples chilled by, but

    FIG. 1. The orf162b locus and flanking regions. The arrow indicates the direction of transcription from a promoter immediately 59 of puhA. The 2-kbp BamHIfragment was cloned in pUC13, and the restriction sites shown were used to disrupt orf162b, construct a complementation plasmid, and produce a probe for RNA blothybridization. The solid horizontal bar shows the extent of this BsaBI-to-EcoRI probe.

    VOL. 182, 2000 R. CAPSULATUS orf162b 5441

    on April 5, 2021 by guest

    http://jb.asm.org/

    Dow

    nloaded from

    http://jb.asm.org/

  • not immersed in, liquid nitrogen. Flash spectroscopy was carried out as previ-ously described (17).

    SDS-PAGE. The amount of protein in each chromatophore preparation wasdetermined by a modified Lowry method, with bovine serum albumin as thestandard (21). Samples containing 40 mg of protein were mixed with loadingbuffer, heated at 50°C for 10 min, and used in a Tricine-SDS-polyacrylamide gelelectrophoresis (PAGE) system (23). Gels were stained in a solution of 0.025%Coomassie brilliant blue G-250 in 40% methanol and 10% acetic acid and weredestained in the same solution lacking the dye.

    RESULTS

    orf162b is transcribed as part of the puhA superoperon.Hybridization of RNA blots with an orf162b-specific probe(Fig. 1) showed that orf162b is transcribed to produce mes-sages at higher levels in oxygen-deprived than in highly aeratedSB1003 cells (Fig. 2); transcripts were also detected in photo-synthetically grown cultures (data not shown). The largestRNA species detected was approximately 2.7 kb in size and wasdistinct from a smear of less than 1.0 kb, indicating that deg-radation of the primary transcript is rapid. The probe alsobound to molecules of 1.5 kb from SBK1 cultures (attributed totranscription originating at the kanamycin resistance cartridgepromoter, which would be complementary to the 39 end of theprobe [Fig. 1]). There was nonspecific hybridization of theprobe to rRNA in all six samples (sizes of 1.6 and 1.0 kb) (31).However, no 2.7-kb species or low-molecular-weight smearswere detected in lanes containing RNA from strain DW1, inwhich transcription is terminated at the V cartridge insertedinto puhA (28). Therefore, orf162b is cotranscribed with thepuhA and orf214 genes and is part of a bchFNBHLM-lhaA-puhA-orf214-orf162b superoperon (4).

    Photosynthetic growth kinetics of the orf162b mutant andcomplemented strains. The partial deletion and KIXX disrup-tion of orf162b in strain SBK1 impaired photosynthetic growth,as evidenced by a progressive slowing of growth and often apremature stationary phase compared to the growth of SB1003(Fig. 3). Complementation of strain SBK1 with the orf162bgene borne on plasmid pAH8 resulted in a photosyntheticgrowth pattern similar to that of the parental strain. Thus itseems that orf162b transcripts are translated, the resultantOrf162b protein is required to sustain rapid photosyntheticgrowth, and the wild-type allele is dominant over the disruptedorf162b allele.

    Evaluation of the presence and function of RC and LHcomplexes by SDS-PAGE and spectroscopy. The levels of RCand LH proteins in chromatophores isolated from oxygen-

    deprived (dark) and anaerobic photosynthetic (illuminated)cultures of strains SBK1 and SB1003 were compared by SDS-PAGE (Fig. 4). In both modes of growth, the orf162b mutantstrains had lower levels of the RC proteins H, M and L, andwas deficient in the LHI a protein. The level of the LHI bprotein could not be ascertained due to its comigration withthe LHII b protein. The levels of the LHII a and b proteinswere not greatly affected by the orf162b mutation.

    Room temperature absorption spectra of SBK1 intact cellsindicated a decrease in the 870-nm LHI shoulder on the850-nm LHII peak (not shown). Absorption spectra at 77 K(liquid N2) of chromatophores isolated from oxygen-deprivedand photosynthetic cultures of SBK1 and SB1003 clearly re-vealed a significant decrease in the LHI 870-nm absorption(shoulder of the LHII 850-nm peak in Fig. 5). These spectrashow that the level of the LHI complex is lower in the orf162bmutant than in the parental strain SB1003, and that the differ-ence is more pronounced in photosynthetically grown cells.The LHII peaks at 800 and 850 nm were marginally increasedby the orf162b mutation in these samples, which were normal-ized on the basis of Bchl content.

    The magnitude of the SBK1 RC Bchl special pair bleachingin response to a train of eight flashes of actinic light was greatlyreduced in this orf162b mutant relative to SB1003 (Fig. 6a).These and other experiments indicate that strain SBK1 con-tains approximately 33% (in oxygen-deprived cells) or 3% (incells from photosynthetically grown cultures in premature sta-tionary phase) of the level of RCs in SB1003 cells grown to thesame cell density under identical conditions, based on theextent of RC bleaching after the eighth flash.

    The carotenoid bandshift of R. capsulatus, like that of R.sphaeroides, is an intrinsic monitor of transmembrane poten-tials that change in response to light-driven electron and pro-ton translocation (15). In concert with the measurements of

    FIG. 2. Blot of R. capsulatus RNA hybridized with the orf162b probe (Fig. 1).RNA was isolated from highly aerated and oxygen-deprived cultures of DW1,SB1003, and SBK1. The approximate sizes of relevant signals (in kilobases) aregiven on the left.

    FIG. 3. Effect of orf162b mutation and trans complementation with pAH8 onthe photosynthetic growth of R. capsulatus strains.

    5442 AKLUJKAR ET AL. J. BACTERIOL.

    on April 5, 2021 by guest

    http://jb.asm.org/

    Dow

    nloaded from

    http://jb.asm.org/

  • the RC, the carotenoid bandshift after a single flash wassmaller in the orf162b-disrupted strain SBK1 than in SB1003,both in the absence (Fig. 6b) and in the presence (Fig. 6c) ofantimycin, which inhibits the cytochrome b/c1 complex (27). Allthree phases of the bandshift were apparent in the single-flashabsorbance changes of the deletion strain in the absence ofantimycin (Fig. 6b), indicating that functional RCs were con-nected appropriately to the cytochrome b/c1 complex. Multipleflashes resulted in successive turnovers of the RC and progres-

    sively greater transmembrane potentials, but to a lesser extentin SBK1 than in SB1003, and the difference was more pro-nounced in chromatophores from cells grown photosyntheti-cally (Fig. 6d).

    In summary, the flash-induced spectral changes indicate thatthe RC of SBK1 cells is present in only small amounts, but thatneither the catalytic activity of the RC nor the quinone transferfrom the RC to the b/c1 complex is impaired. The spectro-scopic data, like the SDS-PAGE data, show that this mutationreduces the amounts of the RC and LHI complexes, and reveala more severe effect of the orf162b disruption after an initialperiod of rapid growth under photosynthetic conditions.

    pufB (LHI b) gene fusion expression in the orf162b mutantstrain SBK1. To evaluate whether the reduction in LHI andRC levels in SBK1 might be due to decreased expression of thepuf genes that encode LHI and two of the three RC proteins,the b-galactosidase activities expressed from a translationallyin-frame pufB::lacZ gene fusion, which is transcribed from thepuf promoter in plasmid pXCA601::935, in strains SBK1(pXCA601::935) and SB1003(pXCA601::935) were determined.During oxygen-deprived growth, the SBK1(pXCA601::935)b-galactosidase activities ranged from 84 to 95% of the SB1003(pXCA601::935) values, and during photosynthetic growth therelative activity was 155%. We conclude that neither initiationof transcription from the puf promoter nor initiation of trans-lation of the LHI b protein is significantly decreased by theorf162b mutation, and that the Orf162b protein functions post-translationally to yield appropriately large amounts of RC andLHI complexes.

    Suppression of the orf162b phenotype. An agar plate spreadwith about 106 SBK1 cells that had reached premature station-ary phase during photosynthetic growth was incubated underphotosynthetic growth conditions, and about 150 coloniesformed within 2 to 4 days. Four strains in which the orf162bphotosynthetic growth phenotype was thus suppressed (desig-nated SBK18, SBK20, SBK21, and SBK23) were isolated fromthese colonies. When grown photosynthetically in liquid cul-ture, these suppressor strains did not exhibit a premature slow-ing of growth, in contrast to the SBK1 parental strain (Fig. 7a).During growth under oxygen-deprived conditions, all four sup-pressor strains were found to contain less of the LHII complex

    FIG. 4. Effects of the orf162b mutation on RC and LHI protein levels inchromatophores isolated from oxygen-deprived or photosynthetic cultures ofSBK1 (orf162b mutant) compared to SB1003 (wild type), as revealed by SDS-PAGE.

    FIG. 5. Low-temperature absorption spectra of chromatophores from SB1003 and SBK1 cultures grown under either oxygen-deprived or photosynthetic condi-tions.

    VOL. 182, 2000 R. CAPSULATUS orf162b 5443

    on April 5, 2021 by guest

    http://jb.asm.org/

    Dow

    nloaded from

    http://jb.asm.org/

  • FIG. 6. Flash spectroscopy analysis of chromatophores from SB1003 and SBK1 strains. (a) The amount of functional RC. (b) Single-flash carotenoid bandshiftwithout antimycin. (c) Single-flash carotenoid bandshift with antimycin. (d) Carotenoid bandshift over eight flashes without antimycin. The vertical bars on the leftrepresent 0.00435 absorbance units at the wavelength pairs indicated.

    5444 AKLUJKAR ET AL. J. BACTERIOL.

    on April 5, 2021 by guest

    http://jb.asm.org/

    Dow

    nloaded from

    http://jb.asm.org/

  • than did SBK1 and the parental strain SB1003, as indicated bydecreases in the 800- and 850-nm peaks of room temperatureabsorption spectra of intact cells (Fig. 7b; for brevity, onlySBK23 data are shown). When grown photosynthetically to the

    density at which SBK1 enters the characteristic prematurestationary phase, SBK18 and SBK21 contained less LHII thanSBK1 and SB1003; the level of LHII in SBK20 and SBK23 wascomparable to that in SB1003 but less than that in SBK1 (Fig.7c). In the stationary phase of photosynthetic growth, all foursuppressor strains had approximately the same level of LHII asSB1003 (not shown). The spectra in Fig. 7b and c also showthat SBK1 contains more LHII than SB1003 when samples arenormalized to equivalent cell numbers. The frequency of sup-pressor mutations (ca. 1.5 3 1024) and the different effects onLHII levels indicate that the orf162b mutant phenotype issuppressed by several types of second-site mutations.

    DISCUSSION

    As noted above, in the five species of purple photosyntheticbacteria for which DNA sequence data are available, the puhAgene is followed by orf214-like and orf162b-like ORFs in thesame transcriptional orientation as puhA. In R. capsulatus,genetic experiments indicated that puhA and orf214 are co-transcribed with at least one 39 gene (28), and similar conclu-sions were reached as a result of puhA disruption experimentson R. sphaeroides (9, 24). Our RNA blots provide biochemicalevidence that synthesis of orf162b messages depends on tran-scription that initiates 59 of puhA from either or both of thepromoters located within the lhaA gene and 59 of the bchFgene. Such a long transcript was not detected (Fig. 2), and sowe suggest that degradation of segments of these primary tran-scripts is so rapid that such long molecules are undetectable inblot hybridizations, by analogy to the R. capsulatus crtEF-bchXYZ-pufQBALMX superoperon (1, 4). This rapid degrada-tion may explain the extremely small amounts of the orf162b2.7-kb RNA segment detected with the orf162b probe, and whyBauer et al. (3) did not detect an orf162b precursor RNAmolecule in blots probed with a puhA-specific probe. Thisinterpretation implies that the bchFNBHLM-lhaA-puhA-orf214 superoperon of R. capsulatus (28) includes orf162b,which suggests a physiological connection among the puhA,orf214, and orf162b genes.

    Hybridization of the orf162b probe to a 1.5-kb transcript inSBK1 suggests that the 39 region of the disrupted orf162b geneis transcribed from the kanamycin resistance cartridge pro-moter. The reason why this transcript was more abundant inhighly aerated SBK1 cultures is not known. This transcriptionmay or may not cause overexpression of downstream ORFs,which have no known function at present. However, such hy-pothetical overexpression could not be responsible for the im-paired photosynthetic growth of strain SBK1, because plasmidpAH8 (which expresses orf162b from the puf promoter) re-stores normal growth, indicating that the Orf162b protein isproduced in wild-type R. capsulatus and is required for optimalphotosynthetic growth.

    Disruption of orf162b leads to impaired photosyntheticgrowth, typically exhibited as a premature stationary phase.Our spectroscopic and SDS-PAGE data demonstrate thatSBK1 is deficient in the RC and LHI proteins. These findingsindicate that the function of Orf162b is to promote the expres-sion, assembly, or stabilization of these pigment-protein com-plexes. The presence or absence of the Orf162b protein did notsignificantly reduce initiation of transcription from the pufpromoter or translation of a pufB::lacZ fusion, and so the effectof Orf162b on the LHI and the RC complexes appears to bemanifested posttranslationally. We propose that Orf162b en-hances the assembly or stability of RC-LHI core complexes.

    The impaired photosynthetic growth of SBK1 was overcomeby frequent spontaneous secondary mutations that led to a

    FIG. 7. (a) Photosynthetic growth of suppressor strains SBK18, SBK20,SBK21, and SBK23, compared to growth of SB1003 and SBK1. (b and c) Ab-sorption spectra (normalized to light scattering at 650 nm) of intact cells fromoxygen-deprived (b) and photosynthetic (c) cultures of the representative sup-pressor strain SBK23 against those of the wild-type SB1003 and orf162b mutantSBK1 strains. The vertical axes give absorbance units, and the horizontal axesgive wavelengths (in nanometers).

    VOL. 182, 2000 R. CAPSULATUS orf162b 5445

    on April 5, 2021 by guest

    http://jb.asm.org/

    Dow

    nloaded from

    http://jb.asm.org/

  • decrease in the amount of LHII. This might indicate that thepremature slowing of growth is due to lower levels of the RCand LHI complexes caused by an increase in the ratio of LHIIto LHI-RC core complexes, rather than a simple decrease inthe amounts of the core complex. It is conceivable that theOrf162b protein provides an optimal arrangement of com-plexes within the ICM of R. capsulatus such that, in the absenceof Orf162b, LHII mixes with the core complex and interfereswith assembly. Alternatively, Orf162B may function to removea growth-inhibiting factor associated with excess LHII. Re-gardless, the orf162b and suppressor mutant phenotypes sug-gest that the function of Orf162b relates to all three complexesof the photosynthetic apparatus. Because an orf162b-like ORFis present in R. rubrum, which lacks LHII, and because we haveobserved an orf162b mutant phenotype in an LHII2 R. capsu-latus background (14a), we suggest that the effects of orf162bdisruption on LHII are secondary to a primary effect on theRC and LHI.

    The core complex deficiencies of SBK1 were more pro-nounced when photosynthetic growth slowed after a brief pe-riod of rapid cell division than when growth of this mutant hadslowed due to oxygen deprivation, and so this low-oxygengrowth condition appears to compensate in some measure forthe loss of Orf162b function. We entertained the possibilitythat the presence of oxygen allows the oxidation of a cofactorrequired for photosynthesis and that Orf162b catalyzes thisoxidation under anaerobic conditions; however, inclusion ofthe alternative electron acceptor dimethyl sulfoxide (14) in themedium of photosynthetic cultures did not compensate fororf162b disruption (unpublished data). We favor an alternativehypothesis, that the low rate of cell division (and ICM synthe-sis) in the pseudostationary phase of oxygen-limited culturesallows the inefficient accumulation of core complexes in theICM, whereas cell division upon transfer to anaerobic, illumi-nated conditions of growth is too rapid for commensurateaccumulation of core complexes, leading to a gradual decreasein the number of core complexes per cell so that rapid growthcannot be sustained. This model will be investigated in futureexperiments.

    We suggest that the ORFs in R. sphaeroides, R. gelatinosus,R. rubrum, and R. palustris that resemble orf162b also encodeproteins which optimize the levels of RC-LHI core complexesand that these ORFs are genes that are cotranscribed withpuhA. If these Orf162b-like sequences found in different spe-cies are indeed homologous, they exhibit remarkable evolu-tionary divergence. A predicted transmembrane domain com-mon to the N-terminal regions of all five sequences suggeststhat membrane attachment and hence the proximity of Orf162bto the membrane-bound photosynthetic apparatus may be im-portant for the accumulation of RC-LHI core complexes in theICM. This function may require Orf162b and related proteinsto associate intimately with other proteins that have evolveddivergently in different species, but in parallel with them withineach species. Further studies of Orf162b function will be ac-companied by analyses of the orf153 locus in R. sphaeroides, theorf154 locus in R. gelatinosus, the orfI3087 locus in R. rubrum,and the orf168 locus in R. palustris to assess whether theseORFs are functionally equivalent to the R. capsulatus orf162bgene.

    In conclusion, our analyses of the orf162b locus revealed thatit encodes a protein that affects the relative levels of the RC,LHI, and LHII complexes and is required for optimal photo-synthetic growth of R. capsulatus. The transcriptional coregu-lation of orf162b with the puhA gene in R. capsulatus and thepresence of ORFs similar to orf162b at the same chromosomallocation relative to puhA in other species suggest that Orf162b-

    like proteins perform a similar function in a variety of purplephotosynthetic bacteria.

    ACKNOWLEDGMENTS

    We thank W. Collins for construction of the pUC13::BamHI plas-mid and C. Harwood, S. Kaplan, and K. Nagashima for provision ofunpublished data.

    M.A. was supported in part by a fellowship from UBC, and thisresearch was funded by NSERC (Canada) grant 5-82796 to J.T.B.

    REFERENCES

    1. Adams, C. W., M. E. Forrest, S. N. Cohen, and J. T. Beatty. 1989. Structuraland functional analysis of transcriptional control of the Rhodobacter capsu-latus puf operon. J. Bacteriol. 171:473–482.

    2. Alberti, M., D. E. Burke, and J. E. Hearst. 1995. Structure and sequence ofthe photosynthetic gene cluster, p. 1083–1106. In R. E. Blankenship, M. T.Madigan, and C. E. Bauer (ed.), Anoxygenic photosynthetic bacteria. KluwerAcademic Publishers, Dordrecht, The Netherlands.

    3. Bauer, C. E., J. Buggy, Z. Yang, and B. L. Marrs. 1991. The superoperonalorganization of genes for pigment biosynthesis and reaction center proteinsis a conserved feature in R. capsulatus: analysis of overlapping bchB andpuhA transcripts. Mol. Gen. Genet. 228:438–444.

    4. Beatty, J. T. 1995. Organization of photosynthesis gene transcripts, p. 1209–1219. In R. E. Blankenship, M. T. Madigan, and C. E. Bauer (ed.), Anoxy-genic photosynthetic bacteria. Kluwer Academic Publishers, Dordrecht, TheNetherlands.

    5. Beatty, J. T., and H. Gest. 1981. Generation of succinyl-coenzyme A inphotosynthetic bacteria. Arch. Microbiol. 129:335–340.

    6. Bérard, J., and G. Gingras. 1991. The puh structural gene coding for the Hsubunit of the Rhodospirillum rubrum photoreaction center. Biochem. Cell.Biol. 69:122–131.

    7. Bibb, M. J., and S. N. Cohen. 1982. Gene expression in Streptomyces: con-struction and application of promoter-probe plasmid vectors in Streptomyceslividans. Mol. Gen. Genet. 187:265–277.

    8. Bollivar, D. W., J. Y. Suzuki, J. T. Beatty, J. M. Dobrowski, and C. E. Bauer.1994. Directed mutational analysis of bacteriochlorophyll a biosynthesis inRhodobacter capsulatus. J. Mol. Biol. 237:622–640.

    9. Chen, X.-Y., V. Yurkov, M. L. Paddock, M. Y. Okamura, and J. T. Beatty.1998. A puhA gene deletion and plasmid complementation system for facilesite directed mutagenesis of the reaction center H protein of Rhodobactersphaeroides. Photosyn. Res. 55:369–373.

    9a.Choudhary, M., and S. Kaplan. 2000. DNA sequence analysis of the pho-tosynthesis region of Rhodobacter sphaeroides 2.4.1T. Nucleic Acids Res.28:862–867.

    10. Claros, M. G., and G. von Heijne. 1995. TopPred II, version 1.1: predictionof transmembrane segments in integral membrane proteins, and the putativetopologies. CABIOS 10:685–686.

    11. Ditta, G., T. Schmidhauser, E. Yakobsen, P. Lu, X.-W. Liang, D. R. Finlay,D. Guiney, and D. R. Helinski. 1985. Plasmids related to the broad hostrange vector, pRK290, useful for gene cloning and for monitoring geneexpression. Plasmid 13:149–153.

    12. Drews, G., and J. R. Golecki. 1995. Structure, molecular organization, andbiosynthesis of membranes of purple bacteria, p. 231–257. In R. E. Blanken-ship, M. T. Madigan, and C. E. Bauer (ed.), Anoxygenic photosyntheticbacteria. Kluwer Academic Publishers, Dordrecht, The Netherlands.

    13. Feick, R., R. van Grondelle, C. P. Rijgersberg, and G. Drews. 1980. Fluo-rescence emission by wild-type and mutant strains of Rhodopseudomonascapsulata. Biochim. Biophys. Acta 593:241–253.

    14. Ferguson, S. J., J. B. Jackson, and A. G. McEwan. 1987. Anaerobic respi-ration in the Rhodospirillaceae: characterization of pathways and evaluationof roles in redox balancing during photosynthesis. FEMS Microbiol. Rev.46:117–143.

    14a.Harmer, A. L. 1998. M. Sc. thesis, University of British Columbia, Vancou-ver, British Columbia, Canada.

    15. Jackson, J. B., and A. R. Crofts. 1969. The high energy state in chromato-phores from Rhodopseudomonas sphaeroides. FEBS Lett. 4:185–189.

    16. LeBlanc, H., A. S. Lang, and J. T. Beatty. 1999. Transcript cleavage, atten-uation, and an internal promoter in the Rhodobacter capsulatus puc operon.J. Bacteriol. 181:4955–4960.

    17. Lilburn, T. G., C. E. Haith, R. C. Prince, and J. T. Beatty. 1992. Pleiotropiceffects of pufX gene deletion on the structure and function of the photosyn-thetic apparatus of Rhodobacter capsulatus. Biochim. Biophys. Acta 1100:160–170.

    18. Manoil, C. 1991. Analysis of membrane protein topology using alkalinephosphatase and b-galactosidase. Methods Cell Biol. 34:61–75.

    19. Messing, J. 1983. New M13 vectors for cloning. Methods Enzymol. 101:20–78.

    20. Papiz, M. Z., S. M. Prince, A. M. Hawthornthwaite-Lawless, G. McDermott,A. A. Freer, N. W. Isaacs, and R. J. Cogdell. 1996. A model for the photo-

    5446 AKLUJKAR ET AL. J. BACTERIOL.

    on April 5, 2021 by guest

    http://jb.asm.org/

    Dow

    nloaded from

    http://jb.asm.org/

  • synthetic apparatus of purple bacteria. Trends Plant Sci. 1:198–206.21. Peterson, G. 1983. Determination of total protein. Methods Enzymol. 91:

    95–119.22. Sambrook, J., E. F. Fritsch, and T. Maniatis. 1989. Molecular cloning: a

    laboratory manual, 2nd ed. Cold Spring Harbor Laboratory Press, ColdSpring Harbor, N.Y.

    23. Schägger, H., and H. von Jagow. 1987. Tricine-sodium dodecyl sulfate-polyacrylamide gel electrophoresis for the separation of proteins in the rangefrom 1 to 100 kDa. Anal. Biochem. 166:368–379.

    24. Sockett, R. E., T. J. Donohue, A. R. Varga, and S. Kaplan. 1989. Control ofphotosynthetic membrane assembly in Rhodobacter sphaeroides mediated bypuhA and flanking sequences. J. Bacteriol. 171:436–446.

    25. Solioz, M., and B. Marrs. 1977. The gene transfer agent of Rhodopseudo-monas capsulata. Arch. Biochem. Biophys. 181:300–307.

    26. Taylor, D. P., S. N. Cohen, W. G. Clark, and B. L. Marrs. 1983. Alignmentof genetic and restriction maps of the photosynthesis region of the Rhodo-pseudomonas capsulata chromosome by a conjugation-mediated marker res-cue technique. J. Bacteriol. 154:580–590.

    27. van den Berg, W. H., R. C. Prince, C. L. Bashford, K. Takamiya, W. D.

    Bonner, and P. L. Dutton. 1979. Electron and proton transport in the ubiqui-none-cytochrome b-c2 oxidoreductase of Rhodopseudomonas sphaeroides:patterns of binding and inhibition by antimycin. J. Biol. Chem. 254:8594–8604.

    28. Wong, D. K.-H., W. J. Collins, A. Harmer, T. G. Lilburn, and J. T. Beatty.1996. Directed mutational analysis of the Rhodobacter capsulatus puhA geneand orf214: pleiotropic effects on photosynthetic reaction center and light-harvesting I complexes. J. Bacteriol. 178:2334–2342.

    29. Yen, H. C., N. T. Hu, and B. L. Marrs. 1979. Characterization of the genetransfer agent made by an overproducer mutant of Rhodopseudomonas cap-sulata. J. Mol. Biol. 131:157–168.

    30. Young, C. S., R. C. Reyes, and J. T. Beatty. 1998. Genetic complementationand kinetic analysis of Rhodobacter capsulatus ORF1696 mutants indicatethat the ORF1696 protein enhances assembly of the light-harvesting I com-plex. J. Bacteriol. 180:1759–1765.

    31. Yu, P.-L., B. Hohn, H. Falk, and G. Drews. 1982. Molecular cloning of theribosomal RNA genes of the photosynthetic bacterium Rhodopseudomonascapsulata. Mol. Gen. Genet. 188:392–398.

    VOL. 182, 2000 R. CAPSULATUS orf162b 5447

    on April 5, 2021 by guest

    http://jb.asm.org/

    Dow

    nloaded from

    http://jb.asm.org/