16
Heme, a Plastid-Derived Regulator of Nuclear Gene Expression in Chlamydomonas W Erika D. von Gromoff, a Ali Alawady, b Linda Meinecke, a Bernhard Grimm, b and Christoph F. Beck a,1 a Fakulta ¨ t fu ¨ r Biologie, Institut fu ¨ r Biologie III, Universita ¨ t Freiburg, D-79104 Freiburg, Germany b Institut fu ¨ r Biologie/Pflanzenphysiologie, Humboldt Universita ¨ t, D-10115 Berlin, Germany To gain insight into the chloroplast-to-nucleus signaling role of tetrapyrroles, Chlamydomonas reinhardtii mutants in the Mg-chelatase that catalyzes the insertion of magnesium into protoporphyrin IX were isolated and characterized. The four mutants lack chlorophyll and show reduced levels of Mg-tetrapyrroles but increased levels of soluble heme. In the mutants, light induction of HSP70A was preserved, although Mg-protoporphyrin IX has been implicated in this induction. In wild-type cells, a shift from dark to light resulted in a transient reduction in heme levels, while the levels of Mg-protoporphyrin IX, its methyl ester, and protoporphyrin IX increased. Hemin feeding to cultures in the dark activated HSP70A. This induction was mediated by the same plastid response element (PRE) in the HSP70A promoter that has been shown to mediate induction by Mg-protoporphyrin IX and light. Other nuclear genes that harbor a PRE in their promoters also were inducible by hemin feeding. Extended incubation with hemin abrogated the competence to induce HSP70A by light or Mg-protoporphyrin IX, indicating that these signals converge on the same pathway. We propose that Mg-protoporphyrin IX and heme may serve as plastid signals that regulate the expression of nuclear genes. INTRODUCTION During evolution of the chloroplast from a cyanobacterial pro- genitor, most of the genes from the endosymbionts were trans- ferred to the nuclear genomes of their hosts (Martin et al., 2002; Leister, 2003). In the case of Chlamydomonas reinhardtii, the chloroplast retained 72 genes (Maul et al., 2002). However, chlo- roplasts are estimated to harbor ;3000 proteins (Abdallah et al., 2000), the vast majority being encoded by the nucleus, translated in the cytosol, and subsequently transported into chloroplasts. Some of these proteins regulate essential steps of gene expres- sion within the chloroplast (Goldschmidt-Clermont, 1998; Leon et al., 1998). In the course of evolution, the plastids acquired new properties, such as the ability to import proteins and to generate retrograde signals that in turn regulate the expression of nuclear genes. Research in recent years has provided evidence for multiple retrograde signaling pathways by which the chloroplast regulates a subset of nuclear genes, most of which encode chloroplast-localized proteins involved in photosynthesis. Among the pathways studied in some detail, one is dependent on products of plastid protein synthesis, for another the signal is singlet oxygen, a third employs chloroplast-generated hydrogen peroxide, a fourth is controlled by the redox state of the photo- synthetic electron transport chain, and a fifth involves anabolic intermediates and possibly proteins of the tetrapyrrole biosyn- thesis pathway (reviewed in Rodermel, 2001; Gray et al., 2003; Pfannschmidt, 2003; Beck, 2005; Nott et al., 2006). One retrograde signaling pathway for which many data have accumulated involves components of tetrapyrrole biosynthesis (reviewed in Strand, 2004; Beck, 2005; Beck and Grimm, 2006; Nott et al., 2006). In all plants, the steps of tetrapyrrole biosyn- thesis up to protoporphyrinogen IX (Protogen) occur only within the chloroplast but are catalyzed by nuclear-encoded enzymes (Beale, 1999; Papenbrock and Grimm, 2001). In higher plants, a fraction of Protogen leaves the chloroplast by unknown routes and enters the mitochondria where it serves as precursor of heme. Protogen in the chloroplast serves as precursor of both heme and chlorophyll (Papenbrock and Grimm, 2001). In Chla- mydomonas, in contrast with higher plants, the enzymes for the conversion of Protogen into protoporphyrin IX (Proto) and the subsequent insertion of Fe 2þ into Proto by ferrochelatase have recently been suggested to be exclusively located within the plastids (van Lis et al., 2005). In this alga, only single genes exist for Protogen oxidase and ferrochelatase whose products have been shown by immunolocalization to localize to chloroplasts (van Lis et al., 2005). Thus, in Chlamydomonas, all heme require- ments of the cell appear to be met by the chloroplast, implying a constant flux of heme to all cellular compartments. Insertion of magnesium into Proto forms Mg-protoporphyrin IX (MgProto) and is catalyzed by Mg-chelatase (CHL), a membrane interacting enzyme composed of three different subunits: CHLD, CHLH, and CHLI (Willows et al., 1996). In norflurazon-treated Arabidopsis thaliana, an increased accumulation of MgProto has been shown to lead to the repression of a nuclear light harvesting chlorophyll a/b binding protein (Lhcb) gene (Strand et al., 2003). Also, in cress (Lapidium sativum) and barley (Hordeum vulgare), the accumulation of Mg-tetrapyrroles correlated with a reduced expression of nuclear photosynthetic genes (Oster et al., 1996; La Rocca et al., 2001). Furthermore, a genetic screen based on the use of norflurazon has allowed the identification of a series of Arabidopsis genome uncoupled (gun) mutants that are modified 1 Address correspondence to [email protected]. The author responsible for distribution of materials integral to the findings presented in this article in accordance with the policy described in the Instructions for Authors (www.plantcell.org) is: Christoph F. Beck ([email protected]). W Online version contains Web-only data. www.plantcell.org/cgi/doi/10.1105/tpc.107.054650 The Plant Cell, Vol. 20: 552–567, March 2008, www.plantcell.org ª 2008 American Society of Plant Biologists Downloaded from https://academic.oup.com/plcell/article/20/3/552/6092166 by guest on 23 June 2021

Heme, a Plastid-Derived Regulator of Nuclear Gene ...b Institut fu¨r Biologie/Pflanzenphysiologie, Humboldt Universita¨t, D-10115 Berlin, Germany To gain insight into the chloroplast-to-nucleus

  • Upload
    others

  • View
    8

  • Download
    0

Embed Size (px)

Citation preview

  • Heme, a Plastid-Derived Regulator of Nuclear GeneExpression in Chlamydomonas W

    Erika D. von Gromoff,a Ali Alawady,b Linda Meinecke,a Bernhard Grimm,b and Christoph F. Becka,1

    a Fakultät für Biologie, Institut für Biologie III, Universität Freiburg, D-79104 Freiburg, Germanyb Institut für Biologie/Pflanzenphysiologie, Humboldt Universität, D-10115 Berlin, Germany

    To gain insight into the chloroplast-to-nucleus signaling role of tetrapyrroles, Chlamydomonas reinhardtii mutants in the

    Mg-chelatase that catalyzes the insertion of magnesium into protoporphyrin IX were isolated and characterized. The four

    mutants lack chlorophyll and show reduced levels of Mg-tetrapyrroles but increased levels of soluble heme. In the mutants,

    light induction of HSP70A was preserved, although Mg-protoporphyrin IX has been implicated in this induction. In wild-type

    cells, a shift from dark to light resulted in a transient reduction in heme levels, while the levels of Mg-protoporphyrin IX, its

    methyl ester, and protoporphyrin IX increased. Hemin feeding to cultures in the dark activated HSP70A. This induction was

    mediated by the same plastid response element (PRE) in the HSP70A promoter that has been shown to mediate induction by

    Mg-protoporphyrin IX and light. Other nuclear genes that harbor a PRE in their promoters also were inducible by hemin

    feeding. Extended incubation with hemin abrogated the competence to induce HSP70A by light or Mg-protoporphyrin IX,

    indicating that these signals converge on the same pathway. We propose that Mg-protoporphyrin IX and heme may serve as

    plastid signals that regulate the expression of nuclear genes.

    INTRODUCTION

    During evolution of the chloroplast from a cyanobacterial pro-

    genitor, most of the genes from the endosymbionts were trans-

    ferred to the nuclear genomes of their hosts (Martin et al., 2002;

    Leister, 2003). In the case of Chlamydomonas reinhardtii, the

    chloroplast retained 72 genes (Maul et al., 2002). However, chlo-

    roplasts are estimated to harbor ;3000 proteins (Abdallah et al.,2000), the vast majority being encoded by the nucleus, translated

    in the cytosol, and subsequently transported into chloroplasts.

    Some of these proteins regulate essential steps of gene expres-

    sion within the chloroplast (Goldschmidt-Clermont, 1998; Leon

    et al., 1998). In the course of evolution, the plastids acquired new

    properties, such as the ability to import proteins and to generate

    retrograde signals that in turn regulate the expression of nuclear

    genes. Research in recent years has provided evidence for

    multiple retrograde signaling pathways by which the chloroplast

    regulates a subset of nuclear genes, most of which encode

    chloroplast-localized proteins involved in photosynthesis. Among

    the pathways studied in some detail, one is dependent on

    products of plastid protein synthesis, for another the signal is

    singlet oxygen, a third employs chloroplast-generated hydrogen

    peroxide, a fourth is controlled by the redox state of the photo-

    synthetic electron transport chain, and a fifth involves anabolic

    intermediates and possibly proteins of the tetrapyrrole biosyn-

    thesis pathway (reviewed in Rodermel, 2001; Gray et al., 2003;

    Pfannschmidt, 2003; Beck, 2005; Nott et al., 2006).

    One retrograde signaling pathway for which many data have

    accumulated involves components of tetrapyrrole biosynthesis

    (reviewed in Strand, 2004; Beck, 2005; Beck and Grimm, 2006;

    Nott et al., 2006). In all plants, the steps of tetrapyrrole biosyn-

    thesis up to protoporphyrinogen IX (Protogen) occur only within

    the chloroplast but are catalyzed by nuclear-encoded enzymes

    (Beale, 1999; Papenbrock and Grimm, 2001). In higher plants, a

    fraction of Protogen leaves the chloroplast by unknown routes

    and enters the mitochondria where it serves as precursor of

    heme. Protogen in the chloroplast serves as precursor of both

    heme and chlorophyll (Papenbrock and Grimm, 2001). In Chla-

    mydomonas, in contrast with higher plants, the enzymes for the

    conversion of Protogen into protoporphyrin IX (Proto) and the

    subsequent insertion of Fe2þ into Proto by ferrochelatase have

    recently been suggested to be exclusively located within the

    plastids (van Lis et al., 2005). In this alga, only single genes exist

    for Protogen oxidase and ferrochelatase whose products have

    been shown by immunolocalization to localize to chloroplasts

    (van Lis et al., 2005). Thus, in Chlamydomonas, all heme require-

    ments of the cell appear to be met by the chloroplast, implying a

    constant flux of heme to all cellular compartments.

    Insertion of magnesium into Proto forms Mg-protoporphyrin IX

    (MgProto) and is catalyzed by Mg-chelatase (CHL), a membrane

    interacting enzyme composed of three different subunits: CHLD,

    CHLH, and CHLI (Willows et al., 1996). In norflurazon-treated

    Arabidopsis thaliana, an increased accumulation of MgProto has

    been shown to lead to the repression of a nuclear light harvesting

    chlorophyll a/b binding protein (Lhcb) gene (Strand et al., 2003).

    Also, in cress (Lapidium sativum) and barley (Hordeum vulgare),

    the accumulation of Mg-tetrapyrroles correlated with a reduced

    expression of nuclear photosynthetic genes (Oster et al., 1996;

    La Rocca et al., 2001). Furthermore, a genetic screen based on

    the use of norflurazon has allowed the identification of a series of

    Arabidopsis genome uncoupled (gun) mutants that are modified

    1 Address correspondence to [email protected] author responsible for distribution of materials integral to thefindings presented in this article in accordance with the policy describedin the Instructions for Authors (www.plantcell.org) is: Christoph F. Beck([email protected]).W Online version contains Web-only data.www.plantcell.org/cgi/doi/10.1105/tpc.107.054650

    The Plant Cell, Vol. 20: 552–567, March 2008, www.plantcell.org ª 2008 American Society of Plant Biologists

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/20/3/552/6092166 by guest on 23 June 2021

  • in chloroplast-to-nucleus signaling (Susek et al., 1993). Four of

    the mutants (gun2-5) were shown to have lesions in genes

    involved in tetrapyrrole synthesis and/or degradation. In gun5,

    which harbors a partially functional mutation of the CHLH gene,

    the steady state level of MgProto in norflurazon-treated plants

    was reduced relative to the wild type, correlating with a reduced

    repression of Lhcb (Mochizuki et al., 2001; Strand et al., 2003).

    The incubation of leaf protoplasts with MgProto, but not with

    porphobilinogen, Proto, or heme, repressed Lhcb expression,

    suggesting a role of MgProto in gene repression. Transgenic

    tobacco (Nicotiana tabacum) plants that exhibit either over- or

    underexpression of CHLM, the gene encoding MgProto methyl

    transferase, contained either decreased or increased pool levels

    of MgProto, and this correlated with either elevated or reduced

    expression of Lhcb and HemA (glutamyl tRNA reductase [GluTR]),

    GSA (glutamate-1-semialdehyde aminotransferase), and CHLH

    (the H subunit of Mg-chelatase) (Alawady and Grimm, 2005). Also,

    confirming previous findings, a recently described Arabidopsis

    mutant that accumulates MgProto due to a mutation in CHLM

    showed repression of Lhcb also in the absence of norflurazon

    (Pontier et al., 2007).

    The regulation of nuclear genes by chloroplast-derived signals

    has been examined by studying the induction of the nuclear gene

    for the chaperone HSP70 in the green alga Chlamydomonas.

    Feeding of MgProto or its methylester (MgProtoMe) to Chlamy-

    domonas induces HSP70A (Kropat et al., 1997). Recently, a

    cis-acting plastid response element (PRE) with qualities of an

    enhancer was identified in the HSP70A promoter. PRE mediates

    MgProto induction of HSP70A; other genes that harbor PRE in

    their promoters were likewise induced by MgProto feeding

    (Vasileuskaya et al., 2004; von Gromoff et al., 2006). In Chlamy-

    domonas, MgProto appears to mediate the light induction of

    HSP70A seen after a shift of cultures from dark to light. Thus, a

    CHLH mutant, which does not produce MgProto due to a mu-

    tation in the gene for the H-subunit of Mg-chelatase prevented

    the light induction of HSP70A, as did conditions abrogating the

    accumulation of MgProto in wild-type strains, such as the ad-

    dition of cycloheximide or cells that are partially differentiated

    toward gametes (pregametes). However, in all cases, feeding of

    MgProto to cultures growing in the dark resulted in HSP70A

    induction, supporting a role of MgProto as a messenger in the

    signaling pathway by which light induces the HSP70A chaperone

    gene (Kropat et al., 1997, 2000). For this induction, the Mg-

    tetrapyrroles may be made available to the cytosol and the

    nucleus in a light-dependent manner (Kropat et al., 2000; Beck,

    2005). MgProto in the cytosol after Norflurazon treatment has

    recently been observed in Arabidopsis using confocal laser

    scanning spectroscopy (Ankele et al., 2007). The observed light

    induction of HSP70A via tetrapyrroles is supported by an analysis

    of mutations in PRE: mutant PREs abolished induction of a fused

    reporter gene by light and MgProto (von Gromoff et al., 2006).

    However, a unified view on the regulatory role of tetrapyrrole

    biosynthesis is not yet in sight. An alternative hypothesis has

    been proposed that variations in Mg-porphyrin levels are sensed

    at the sites of Mg-porphyrin synthesis by an enzyme complex and

    subsequently transmitted to the cytosol/nucleus (Alawady and

    Grimm, 2005; Beck and Grimm, 2006).

    A new set of mutants affecting genes of CHL subunits and thus

    causing diminished formation of MgProto allowed us to take a

    new look at the role of tetrapyrroles in chloroplast-to-nucleus

    communication. These mutants accumulate heme, which could

    be shown to control the same set of genes previously identified as

    inducible by MgProto. The signaling pathways of both tetrapyr-

    roles converge at the same cis-acting element.

    RESULTS

    Identification of Mutants Defective in CHL Subunits D and H

    To examine the role of MgProto in chloroplast-to-nucleus com-

    munication, we first isolated mutants in the subunits of CHL. To

    isolate mutants, we screened UV-mutagenized lines for chloro-

    phyll deficiency. Four mutants, which were generated by UV

    mutagenesis, lacked chlorophyll and showed the formation of

    yellow-brown colonies. Based on the initial determination of

    Proto levels, these mutants were selected as potentially defec-

    tive in CHL activity. CHL is a heterotrimeric complex composed

    of subunits named D, H, and I (reviewed in Walker and Willows,

    1997). Analysis of the Chlamydomonas genome sequence re-

    vealed a single gene for subunit D, one gene and a presumed

    pseudogene for subunit H, and two genes for subunit I (Lohr

    et al., 2005). To identify the mutant loci, we first identified BAC

    clones containing the functional genes for the various CHL sub-

    units using gene-specific probes. DNA of these BAC clones was

    employed in complementation assays: the BAC DNA was used

    to transform each mutant line; clones that could complement a

    mutant produced colonies capable of autotrophic growth. Two of

    the four mutants were complemented by BAC clones containing

    the CHLD gene (mutants named chlD-1 and chlD-2), and two

    mutants were complemented by BAC clones with CHLH (mu-

    tants named chlH-1 and chlH-2) (Table 1). These data from

    complementation tests with BAC clones were subsequently

    Table 1. Complementation of Chlorophyll-Deficient Mutants with BAC Clones That Harbor the CHL Genes Indicated

    No. of Photoautotrophic Clones Observed after Transformation Using BAC Clones with Genesa

    Mutants Tested CHLH CHLI1 CHLI2 CHLD No DNA

    CF194 chlH-1 16 0 0 0 1

    CF215 chlD-1 0 0 0 16 0

    CF225 chlD-2 0 1 0 134 2

    CF227 chlH-2 14 0 0 0 0

    a Transformations were performed as described in Methods.

    Heme Regulates Nuclear Genes 553

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/20/3/552/6092166 by guest on 23 June 2021

  • confirmed by transformation with the subcloned genes (data not

    shown).

    Characterization of Defects in Mutants chlD-1 and chlD-2

    Chlamydomonas mutants defective in CHLH have previously

    been characterized (Chekounova et al., 2001). We therefore

    focused our mutant characterization on chlD-1 and chlD-2. The

    structure of the Chlamydomonas CHLD gene, as deduced from

    combined genomic and EST sequence data (http://genome.jgi-

    psf.org/Chlre3/Chlre3.home.html), is presented in Figure 1A. The

    mutations in the gene were identified by sequencing of PCR

    fragments generated with genomic DNA of the mutants. In the

    chlD-2 mutant, a transversion mutation was discovered in exon

    12 (a C-to-A change), which resulted in the generation of a stop

    codon (Figure 1A). Analysis of mutant chlD-1 revealed a G-to-T

    transversion at a highly conserved position at the 59 intron side ofthe border between exon 11 and intron 11 (Figure 1A). To

    elucidate the effect of this mutation, the mRNA in this mutant

    was analyzed by RT-PCR employing primers homologous to

    exon 10 and exon 12. RT-PCR of mRNA from chlD-1 yielded two

    products: one slightly smaller than the wild-type product (610

    versus 630 bp) and the other distinctly larger (798 bp).

    The sequence of the short product revealed that removal of

    intron 11 had taken place. But, in addition, 20 nucleotides from

    the 39 end of exon 11 were missing (Figure 1B). The splicingmachinery apparently recognized an alternative 59 splice site inthe sequence GGT//GTA located within exon 11, 20 nucleotides

    upstream from the wild-type splice site. Consensus sequences

    for 59 and 39 splice sites in Chlamydomonas are (C/A)(A/C)GA//GT(G/A) (Silflow, 1998). At the exon side, specificity appears to

    be relaxed since all nucleotides have been observed at these

    three positions (Silflow, 1998). Splicing at this alternative position

    results in a shift of the reading frame that ends, after codons for

    40 amino acids, with a stop codon. In the case of the longer RT-

    PCR product, intron 11 was not removed. In this case, a stop

    codon occurs after 12 intron-derived codons (Figure 1B).

    Proto and Heme Accumulate in CHL Mutants

    While chlorophyll in all four mutants was very low or not detect-

    able, the concentration of noncovalently bound heme in the

    mutants was elevated twofold to fivefold compared with the wild

    type, and the substrate of CHL, Proto, accumulated to levels

    more than ninefold higher than in the wild type (Table 2). The

    levels of MgProto were distinctly lower in all mutants. Mg-

    protoporphyrin IX monomethyl-ester (MgProtoMe) levels also

    were reduced (Table 2). The elevated heme levels suggest that

    the impairment of chlorophyll synthesis in CHL-defective mu-

    tants, resulting in an accumulation of Proto, enhances heme

    synthesis.

    Regulatory Properties of the CHL Mutants

    Since subunits of CHL have been suggested to participate in the

    regulation of nuclear genes (Papenbrock et al., 2000a, 2000b;

    Mochizuki et al., 2001), we first tested whether the light regulation

    of all four functional CHL genes was affected in the chlD and chlH

    mutants. In the parental strain, all four genes exhibited a light-

    induced accumulation of their transcripts (Figure 2). This induc-

    tion was dependent on cytoplasmic protein synthesis since

    presence of the inhibitor cycloheximide prevented this light-

    induced mRNA accumulation (see Supplemental Figure 1 online).

    By contrast, inhibition of plastid protein synthesis by chloram-

    phenicol had no effect (see Supplemental Figure 1 online). The

    chlD and chlH mutants showed accumulation of their mRNAs

    that were similar to those of the wild type (Figure 2). Analysis of

    the expression patterns of the various CHL genes in the chlD and

    chlH mutants revealed no significant changes in light-induced

    mRNA accumulation (Figure 2). Although the kinetics of mRNA

    accumulation exhibited some strain-dependent variations, an

    increase in mRNA levels was observed for all mutants in inde-

    pendent experiments, indicating that light induction is not

    blocked by the accumulation of Proto or reduced levels of Mg-

    tetrapyrroles in the first 4 h of low-level irradiation.

    Figure 1. Structure of the CHLD Gene and Alterations Caused by the

    Mutations chlD-1 and chlD-2.

    (A) Gene CHLD with 12 exons. Dark-gray boxes indicate deduced

    protein coding exons, open boxes indicate introns, and black bars

    indicate 59 and 39 untranslated regions. The positions and sequence

    alterations of mutations chlD-1 and chlD-2 are indicated.

    (B) RT-PCR sequences covering the splice site alteration caused by

    mutation chlD-1 document the changes at the mRNA level. Sequences

    from the wild type and the two chlD-1 mutant products are shown.

    Vertical arrows indicate the end of exon 11. The altered base at the

    junction of exon 11 and intron 11 is indicated by a bold letter in the long

    RT-PCR fragment. Deduced amino acid sequences that differ from the

    wild type are shown in bold and italics. Asterisks indicate stop codons.

    554 The Plant Cell

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/20/3/552/6092166 by guest on 23 June 2021

  • We analyzed the expression of HSP70A in response to light

    and observed an induction in the wild-type strain and in the four

    mutants defective in CHL (Figure 3). A reduction of approximately

    twofold in induced HSP70A mRNA levels compared with the wild

    type was observed in the chlD mutants. Also, progeny of chlD

    mutants from crosses with the wild type showed this reduction,

    indicating that this reduction is apparently not due to second-site

    mutations in the original mutants. We included a previously char-

    acterized CHLH-defective mutant, brs-1 (though a clone of dif-

    ferent origin than the one originally used by Kropat et al., 1997),

    and observed light induction of the heat shock gene. Using the

    four newly isolated mutants, these data document that a block in

    the synthesis of Mg-tetrapyrroles does not prevent the induction

    of HSP70A by light.

    A Role for Heme as an Inducer of HSP70A

    We previously postulated an essential role for MgProto and/or

    MgProtoMe in the light induction of HSP70A (Kropat et al., 1997,

    2000; Beck, 2005). Such a model was supported by the discov-

    ery of sequence elements within the HSP70A promoter called

    PREs that mediate HSP70A induction in response to MgProto

    feeding (von Gromoff et al., 2006). Since in the CHL mutants

    heme was shown to accumulate (Table 2) and heme previously

    has been identified as a regulator of the nuclear gene HEMA

    (Vasileuskaya et al., 2005), we speculated that heme also plays a

    role in HSP70A expression. This was tested by feeding of hemin

    (which harbors a Fe3þ atom, conferring higher stability in solution

    than the Fe2þ contained in heme) to cultures in the dark. The data

    shown in Figure 4A reveal that hemin feeding in the dark induced

    HSP70A mRNA accumulation. The kinetics of mRNA accumu-

    lation upon hemin feeding were delayed compared with MgProto

    feeding, but the degree of induction with 9 mM hemin was similar

    to that seen after MgProto feeding (9 mM) or a shift from dark to

    light (DLS). We conclude that heme, just like MgProto, may

    activate HSP70A, while Proto and protochlorophyllide, as shown

    previously (Kropat et al., 1997), have no inducing effect.

    To verify and extend these studies, we analyzed the induction

    of HSP70A by various concentrations of hemin. As shown in

    Figure 4B, HSP70A mRNA levels rose with an increase in the

    hemin concentration applied to cell cultures in the dark. Maximal

    induction was observed upon feeding of 16 mM hemin. In sub-

    sequent experiments, a final concentration of 9 mM hemin was

    employed.

    Changes in Tetrapyrrole Levels after a Shift from

    Dark to Light

    Exposure of dark-adapted wild-type cells to light resulted in an

    immediate rise (within

  • drop in pool levels in the 20 min samples, followed by a second

    increase. Such fluctuations, also observed by others (Pöpperl

    et al., 1998), may reflect a fine-tuning process among the enzy-

    matic activities of CHL and downstream enzymes affecting the

    regulatory equilibrium after a sudden shift of cultures from dark to

    light. Upon extended exposure of cultures to light (i.e., >60 min),

    the increase of tetrapyrroles stops and reduced pool levels were

    observed for Proto and MgProto (Figure 5). A similar rapid and

    transient, light-induced increase in pool sizes of MgProto and

    MgProtoMe has also been observed in higher plants grown in

    light/dark cycles (Pöpperl et al., 1998; Papenbrock et al., 1999).

    By contrast, levels of total noncovalently bound heme, deter-

    mined from the same cell cultures, exhibited an approximately

    twofold transient reduction in pool levels after a delay of ;10 min(Figure 5). This decrease may possibly be accounted for by an

    enhanced demand of heme for assembly into proteins. Alterna-

    tively, the decrease in heme levels may reflect a channeling of

    Proto into the Mg-branch for chlorophyll biosynthesis. A prefer-

    ential channeling may be explained by an activation of CHL under

    these conditions. Indeed, an increase in CHL activity upon shift of

    cells from dark to light was observed (Figure 6). This increase

    showed some fluctuations (Figure 6). An elevated CHL activity

    seen 5 min after start of illumination was followed by a decrease

    seen at 10 min, and this again was followed by an increase.

    These initial fluctuations in CHL activity are mirrored by changes

    in MgProto and MgProtoMe pool levels right after start of

    illumination (Figure 5). In the latter, an initial peak at 10 min was

    followed by lowered pool levels at 20 min. However, whether

    there is a strict interdependence between the fluctuations at the

    level of enzyme activity and pool levels remains to be investi-

    gated. The rapid change in CHL activity precludes an enhanced

    de novo synthesis of the enzyme as an explanation for the

    increase in activity. Rather, the data support a light activation of

    the enzyme at the posttranslational level, possibly via thiore-

    doxins, which have been shown to interact with CHL. A stimu-

    lation of the enzyme by reduction was observed in isolated

    chloroplasts (Balmer et al., 2003; Ikegami et al., 2007).

    We also analyzed the tetrapyrrole pools of the mutants upon

    shift from dark to light. No significant changes were observed

    (i.e., heme and Proto pools remained high; Table 2), while the

    Mg-tetrapyrrole pools remained at a low level (see Supplemental

    Figure 2 online).

    Is the Same cis Sequence Element Involved in the Induction

    of HSP70A by Heme and MgProto?

    The data presented above suggest that heme and the magne-

    sium containing porphyrins act through a similar or even identical

    signaling pathway for the induction of HSP70A. To study this is-

    sue in more detail, we tested whether heme induction is possibly

    mediated via the same regulatory regions within the HSP70A

    promoter. For these analyses, we used constructs in which the

    HSP70A promoter (PA), fused to the RBCS2 promoter (PR), was

    located upstream of a reporter gene (HSP70B-TAG) (Figure 7A).

    In transformants with such constructs, induction of the reporter

    gene by light or MgProto has previously been shown to occur via

    activation of the PR promoter by enhancer elements within PA(von Gromoff et al., 2006). Thirty transformants with each of the

    constructs shown in Figure 7A were analyzed for expression of

    the reporter gene after shift from dark to light (DLS) or after

    feeding (in the dark) of MgProto or hemin (9 mM final concentra-

    tion each). In agreement with previous data, MgProto activation

    Figure 3. Light Induction of HSP70A in the Wild Type and Mutants Defective in the D and H Subunits of CHL.

    Cultures grown in the dark at time 0 were shifted into dim light (10 mE m�2 s�1), and samples for RNA isolation were taken at the time points indicated

    (hours). RNA gel blot hybridizations were performed using probes specific for HSP70A and CBLP, the latter serving as a loading control. The quantitative

    evaluation of the RNA gel blots in the case of the wild type and mutants chlH-1, chlH-2, brs-1, comprised four independent experiments. In the case of

    chlD-1 and chlD-2, the data from three independent experiments with the original mutant strains and of four mutant progenies from crosses with the wild

    type were combined. The average levels of mRNA accumulation relative to the dark control (corrected for differences in loading) 6 SE are given.

    556 The Plant Cell

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/20/3/552/6092166 by guest on 23 June 2021

  • of PR was seen in transformants even with the shortest HSP70A

    promoter fragment (Figure 7A). This fragment, named regulatory

    region I, has been shown to harbor an enhancer element that may

    confer inducibility by MgProto to heterologous promoters (von

    Gromoff et al., 2006). Clearly, heme activation also is mediated

    via regulatory region I (Figure 7A).

    We next tested whether the element that mediates induction by

    MgProto within regulatory region I also is involved in the activa-

    tion by hemin. Within regulatory region I, the PRE sequence motif

    GCGACNAN15TA was shown to confer MgProto responsiveness

    (von Gromoff et al., 2006). Two mutant constructs that lacked

    sequences either from the 59 or the 39 ends of regulatory region I(mutations 1 and 2 in Figure 7B), and thus were defective in PRE,

    were tested for transcriptional activation of the reporter gene.

    Transformants with these constructs did not show a response to

    either hemin or MgProto. By contrast, transformants with the

    wild-type construct or a deletion construct lacking five internal

    base pairs but maintaining the PRE motif (mutation 3) were

    inducible by both tetrapyrroles (Figure 7B). From these results we

    conclude that heme induction appears to be mediated via PREs,

    suggesting also that it may employ the same or similar protein

    components as MgProto to activate HSP70A.

    Hemin Activates Other MgProto-Responsive Genes

    We have previously identified a group of nuclear genes that were

    inducible by MgProto feeding (Vasileuskaya et al., 2004; von

    Gromoff et al., 2006). All of these genes harbor in their promoter

    regions the cis-acting PRE motif (von Gromoff et al., 2006). The

    response of these genes to MgProto in wild-type strain CC-124

    was confirmed (Figure 8). All five genes also responded to hemin

    feeding, although the kinetics and the degree of mRNA accumu-

    lation showed some variation compared with MgProto feeding,

    suggesting the involvement of additional factors in the expres-

    sion of these genes. The effect of the cell wall on the kinetics of

    hemin induction was assessed using the cell wall–deficient mu-

    tant CW15. The response of the genes in this mutant to hemin

    appears to be more rapid for all genes (Figure 8), suggesting that

    the cell wall represents some barrier for hemin uptake. These

    results show that in Chlamydomonas a number of genes exists

    that responds to two different tetrapyrroles, heme, and MgProto.

    These data support the concept of an overlapping role for heme

    and MgProto that, as organelle-derived molecules, appear to

    play a regulatory role in transcription within the nucleus.

    Assay of Other Metalated Porphyrins

    To test whether only specific tetrapyrroles play a role in gene reg-

    ulation, we tested the effect of three additional metalated por-

    phyrins. In contrast with the Mg- and Fe-derivatives, neither Zn,

    Co, nor Mn-protoporphyrin IX, when fed to cultures in the dark,

    resulted in induction of HSP70A or HEMA (Figure 9A). Also, Proto

    applied at higher concentrations had no effect on the expression

    of the genes tested (Figure 9B). We previously showed that Proto

    fed to cultures in the dark results in the accumulation of MgProto.

    This MgProto, produced by chloroplast-localized enzymes, ap-

    parently is not accessible to cytosol/nucleus in dark-grown cul-

    tures (Kropat et al., 2000). These results suggest a high degree of

    Figure 4. Changes in HSP70A mRNA Levels after Treatment of Cultures

    in the Dark with Hemin.

    (A) RNA accumulation in wild-type cultures that, after an incubation for

    16 h in the dark (D), were treated with hemin (9 mM), MgProto (9 mM), or

    shifted to light (60 mE m�2 s�1, DLS) for the given number of hours.

    Samples for RNA isolation were taken at the times indicated and RNA gel

    blot hybridizations were performed using probes specific for HSP70A

    and CBLP, the latter serving as a loading control.

    (B) Dependence of HSP70A mRNA accumulation on hemin concentration.

    Cells were harvested, and RNA was extracted 2 h after addition of hemin at

    the concentrations indicated to wild-type cultures incubated in the dark.

    The HSP70A mRNA accumulation observed in a representative experiment

    is given relative to that of the culture that did not receive hemin.

    Heme Regulates Nuclear Genes 557

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/20/3/552/6092166 by guest on 23 June 2021

  • specificity in the interaction of Mg- and Fe-porphyrins with down-

    stream signaling components.

    Extended Incubation with Hemin Results in Turning off the

    Downstream Tetrapyrrole Signaling Pathway

    The incubation of dark-grown cells with hemin resulted in a

    transient HSP70A mRNA accumulation. Twelve hours after the

    addition of hemin, HSP70A mRNA levels were about the same as

    prior to treatment (Figure 10), although hemin was still present at

    ;50% of the initial concentration (data not shown). When such ahemin-treated culture was exposed to light (DLS), at most a very

    weak accumulation of HSP70A mRNA was observed (Figure 10).

    For an explanation we considered the possibility that hemin still

    present in the culture medium may have prevented a light-

    induced accumulation of the Mg-tetrapyrroles (e.g., via feedback

    inhibition of GluTR, the first enzyme specific for tetrapyrrole bio-

    synthesis). Such an accumulation of Mg-tetrapyrroles has pre-

    viously been shown to be essential for the HSP70A activation

    that follows a shift from dark to light (Kropat et al., 2000). An

    increase in Proto, MgProto, and MgProtoMe pool levels indeed

    was not observed when cultures treated with hemin for 12 h in the

    dark were exposed to light (see Supplemental Figure 3 online).

    The defect in light induction in these cells appears to involve

    additional factors. To test whether downstream signaling com-

    ponents may have been affected by hemin treatment, the cul-

    tures were fed with MgProto after 12 h with hemin. No induction

    Figure 6. Changes in CHL Activity after a Shift of Cultures from Dark to

    Light.

    Activity of CHL was determined with cultures of wild-type strain 4Aþgrown in TAP medium to a density of 3 to 5 3 106 cells/mL. Exponentially

    growing cultures after an incubation in dark for 16 to 18 h were

    transferred into light (50 mE m�2 s�1) at time 0. The cells were harvested

    at the time points indicated and disrupted by sonication at 48C. The

    enzyme assays were performed as described in Methods. Each time

    point represents the average of data from two independent experiments.

    In each of these experiments, every time point represents a kinetic

    analysis of MgProto formation from Proto (for details, see Methods).

    Figure 5. Changes in Tetrapyrrole Pool Levels in the Wild Type after a

    Shift from Dark to Light.

    Cultures of the wild type were grown in TAP medium. Prior to dark-to-

    light shift, exponentially growing cultures were incubated in the dark for

    16 to 18 h. At time 0, the cultures were shifted into the light (50 mE m�2

    s�1) and samples were taken at the times indicated. For heme determi-

    nations, 250 mL of cells (4 3 106 cells per mL) were chilled on ice and

    harvested by centrifugation at 48C. For determination of the concentra-

    tion of the other tetrapyrroles, 100 mL of cells (4 3 106 cells per mL) were

    chilled on ice and harvested. The harvested cells quickly were frozen in

    liquid nitrogen and stored at �808C. The samples were processed forpool determinations as outlined in Methods. The average pool levels of

    four to six experiments 6 SE are given.

    558 The Plant Cell

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/20/3/552/6092166 by guest on 23 June 2021

  • was observed (Figure 10), suggesting that either the downstream

    signaling pathway has been turned off or, potentially, hemin

    treatment had activated nucleases that rapidly degrade HSP70A

    mRNA. To address this latter point, heat shock treatment was

    performed with the cells incubated with hemin for 12 h. This

    treatment elicited an initial HSP70A mRNA accumulation similar

    to that observed upon the application of heat shock to hemin-

    free cultures (von Gromoff et al., 1989). We conclude that hemin

    treatment apparently had not impaired the potential of cells to

    accumulate HSP70A mRNA (Figure 10).

    Figure 7. Characterization of HSP70A Promoter Elements That Confer Inducibility by Hemin on the RBCS2 Promoter.

    (A) Expression of the HSP70B-TAG reporter gene with the very weakly expressed Chlamydomonas RBCS2 promoter (PR) after upstream insertion of

    various HSP70A promoter sequences (von Gromoff et al., 2006). The open bar represents HSP70A promoter (PA) sequences. The RBCS2 promoter (PR)

    is represented by a black bar, and the HSP70B reporter gene is indicated by a hatched bar. Transformants harboring the various constructs were

    generated using strain 325 as described previously (von Gromoff et al., 2006). They were either kept in the dark (CD) for 16 to 20 h, exposed to light for

    1 h after the dark period (DLS), treated with MgProto (9 mM) in the dark (MP), or treated with hemin (9 mM) in the dark (H). The reporter gene transcripts

    were detected with the 199-bp TAG probe (Kropat et al., 1995). For each construct, 30 individual transformants were assayed. The number of

    transformants showing induction by the various treatments is indicated. The average levels of RNA accumulation relative to the dark control (corrected

    for differences in loading) 6 SE are given below these numbers. TSA1 and TSA2 designate the transcription start sites used upon heat shock and DLS or

    MgProto treatment in the native HSP70A promoter, respectively. In the constructs shown, the transcription start site employed after shift from dark to

    light or after MgProto treatment is that of the RBCS2 promoter PR (von Gromoff et al., 2006), which is indicated by the boxed designation DLS/MP/H.

    HSE, heat shock element.

    (B) Analysis of the effect of mutations in region I, the �81 to �55 PRE-containing fragment, on the activation of PR by hemin and MgProto. The basicconstruct employed is the third from top in (A). RNA dot blot analyses of 30 transformants for each construct are presented. The deletions (D) introduced

    in region I are indicated. The actual nucleotide sequence located in front of PR is also shown. Shadowed in gray is the consensus sequence of the PRE

    defined previously (von Gromoff et al., 2006). The dot blot signals resulting from hybridization with the TAG probe from individual transformants either

    incubated in the dark, or with 9 mM MgProto, or with 9 mM hemin for 1 h were compared. The fold change in average signal intensity compared with

    cultures kept in the dark 6 SE are given. The last column gives the percentage of transformants that exhibited a signal by dot blot analysis.

    Heme Regulates Nuclear Genes 559

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/20/3/552/6092166 by guest on 23 June 2021

  • DISCUSSION

    Here, we report on experiments that allow a new look at the role

    of tetrapyrroles in chloroplast to nucleus signaling. They are

    based on four newly identified mutants with defects in CHL, the

    first enzyme specific for the chlorophyll biosynthetic branch. Two

    of the mutants were shown to be defective in the H subunit of the

    enzyme since they could be complemented by a wild-type CHLH

    gene (Table 1). Two other mutants are defective in the D subunit.

    These four mutants exhibited either no or only residual levels of

    chlorophyll (Table 2), indicating that the mutations blocked

    chlorophyll synthesis. All four mutants accumulated Proto (Table

    2). In contrast with Chlamydomonas, an accumulation of Proto

    has not been observed in higher plants with deficient CHL activity

    (Papenbrock et al., 2000b). However, accumulation of Proto was

    seen in other Chlamydomonas mutants defective in CHLH (Wang

    et al., 1974; Kropat et al., 2000; Chekounova et al., 2001).

    Analysis of the expression of nuclear genes upon shift of

    cultures from dark to light revealed that the four mutants did not

    exhibit significant differences in the light-induced expression of

    the CHL genes (Figure 2), HEMA, ALAD, or several LHC genes.

    Also, basal level expression of these genes appeared not to be

    affected (L. Meinecke and C.F. Beck, unpublished data).

    The observation that HSP70A was light inducible in the mu-

    tants (Figure 3) at first glance was surprising, since previously a

    mutant defective in CHLH (brs-1) did not show light induction of

    HSP70A (Kropat et al., 1997). A strain with this mutation, al-

    though a different subclone, was included in our studies and here

    showed light induction of HSP70A (Figure 3). The lack of light

    induction in previous experiments may be explained by a sec-

    ond-site mutation in the brs-1 mutant clone originally tested.

    Indeed, we discovered second-site mutations that caused al-

    tered regulation of nuclear genes with high frequency in a screen

    of mutants defective in chlorophyll synthesis. Thus, five out of 13

    Figure 8. Test of MgProto-Inducible Genes for Induction by Hemin.

    Wild-type cells and cells of the cell wall–deficient mutant CW15 were

    incubated in the dark for 16 h and then treated with hemin (9 mM) or

    MgProto (9 mM). Samples for RNA isolation were taken at the times

    indicated. RNA gel blot hybridizations were performed as described in

    Methods using the gene-specific probes described either in Methods or

    by Vasileuskaya et al. (2004). CBLP (von Kampen et al., 1994) served as a

    loading control.

    Figure 9. Test for Inducing Potential of Various Metalated Porphyrins and of Proto at Different Concentrations.

    Cells of the wild type grown in TAP medium were incubated in the dark for 16 to 18 h. At time 0, the various porphyrins were added. Samples for RNA

    isolation were taken at the times indicated. RNA gel blot hybridizations were performed using the probes described in Methods. CBLP served as a

    loading control.

    (A) The metalated porphyrins listed were added to a final concentration of 9 mM.

    (B) The concentrations of Proto indicated were tested for inducing potential.

    560 The Plant Cell

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/20/3/552/6092166 by guest on 23 June 2021

  • chlH mutants from the initial screen for chlorophyll-deficient

    mutants exhibited deregulation of HSP70A. In each case,

    crosses revealed a segregation of the phenotypes of chlorophyll

    deficiency and deregulated gene expression. For our studies, we

    chose two of the chlH mutants that exhibited normal regulation of

    HSP70A. In the two chlD mutants, a reproducible reduction in

    HSP70A mRNA accumulation by a factor of about two was

    observed upon dark-to-light shift compared with the wild type or

    other CHL mutants (Figure 3). In this case, second-site mutations

    being responsible for the reduced induction with high probability

    could be ruled out since all mutant progenies from crosses of the

    two strains with the wild type revealed a similar reduction (data

    not shown). These data point to a complex role of individual CHL

    subunits in signaling, a view supported by the observation that

    Arabidopsis mutants with lesions in the CHLI gene do not exhibit

    a gun phenotype upon norflurazon treatment (Mochizuki et al.,

    2001).

    In view of the accumulated evidence for a crucial role of

    tetrapyrroles in mediating the light induction of HSP70A (Kropat

    et al., 1997, 2000; von Gromoff et al., 2006), we considered an

    involvement of other tetrapyrroles besides the Mg-tetrapyrroles

    in this induction. Heme was an attractive candidate since this

    tetrapyrrole activates HEMA, encoding GluTR, the first enzyme

    specific for tetrapyrrole biosynthesis (Vasileuskaya et al., 2004).

    Indeed, feeding experiments with hemin in the dark induced

    HSP70A mRNA accumulation (Figure 4). This suggested that

    heme may play a similar role to MgProto in signaling.

    A crucial aspect of an understanding of tetrapyrrole gene

    regulation was the definition of a sequence motif within the

    HSP70A promoter that may activate the gene upon hemin

    feeding. The previously identified motif (named PRE) that medi-

    ates HSP70A induction by MgProto (von Gromoff et al., 2006)

    was shown also to confer heme-responsive activation on

    HSP70A since mutations in PRE that abolished induction by

    MgProto also abolished activation by hemin (Figure 7). These

    data provide solid evidence for a specific role of heme in the

    regulation of nuclear genes. Indeed, other genes that harbor a

    PRE were also induced by the feeding in the dark of MgProto or

    hemin (Figure 8). A test of different metalated porphyrins re-

    vealed that only MgProto and hemin induce PRE-harboring

    genes (Figure 9). Previously we have shown that, besides

    MgProto, MgProtoMe2 (used instead of the naturally occurring

    monomethyl ester) also induced HSP70A, but protochlorophyl-

    lide or chlorophyllide had no inducing potential (Kropat et al.,

    1997). This and the inability of metalated porphyrins with metal

    ions other than Mg and Fe to induce HSP70A (Figure 9) point to a

    high degree of specificity in the interaction between the meta-

    lated tetrapyrroles and their receptors. Whether a single receptor

    with dual specificity or two receptors are involved remains to be

    elucidated.

    Salient features of our previous model were two prerequisites

    for HSP70A light activation: first, a light-induced transient accu-

    mulation of Mg-tetrapyrroles and, second, a light-activated

    export of tetrapyrroles from the chloroplast (Kropat et al., 2000;

    Beck, 2005; Beck and Grimm, 2006). Here, in view of our new

    results, we want to modify and extend this model (Figure 11).

    The reduction in heme levels that follows a shift of cultures

    from dark to light (Figure 5) provides new information on

    Figure 10. Response of HSP70A to Various Treatments after Extended

    Incubation with 9 mM Hemin.

    Cultures grown in the dark for 16 h were treated with 9 mM hemin (H) in the

    dark and incubation was continued for 12h (top panel). After this incubation,

    an aliquot of the culture was irradiated (60 mE m�2 s�1) for up to 2 h (DLS,

    second panel from top). To another aliquot, MgProto (MP) was added to a

    final concentration of 9 mM and incubation in the dark was continued for 2 h

    (third panel from top). To a third aliquot, hemin (H) was added (the newly

    added H corresponded to a concentration of 9 mM, but residual hemin was

    present from the previous incubation) and incubation was continued for 2 h

    in the dark (fourth panel from top). In another aliquot, the response of cells to

    heat shock after 12 h incubation with hemin was tested. For this purpose the

    culture was shifted from 23 to 408C in the dark for up to 2 h (fifth panel from

    top). The disappearance of HSP70A mRNA and 18s,25s rRNA 2 h after heat

    shock in cultures treated with hemin has been observed in independent

    experiments, suggesting a destabilization of RNA after prolonged incuba-

    tion at elevated temperature. A dark-to-light shift (DLS) was performed with

    a culture not treated with hemin to illustrate the responsiveness of HSP70A

    to light (sixth panel from top). In all cases, samples for RNA isolation were

    taken at the times indicated. RNA gel blot hybridizations were performed

    using specific probes for HSP70A mRNA and 18s,25s rRNA (labeled 18s

    rRNA), the latter serving as a loading control. In these experiments, CBLP

    could not be used since the mRNA of this gene disappeared upon extended

    hemin treatment (>6 h).

    Heme Regulates Nuclear Genes 561

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/20/3/552/6092166 by guest on 23 June 2021

  • intraplastidal regulatory events that result in the transient accu-

    mulation of Mg-tetrapyrroles upon dark-to-light shift. Three

    different regulatory events may account for these changes. (1)

    In Chlamydomonas cultures, a shift from dark to light not only

    results in a transient increase in Proto and MgProto/MgProtoMe

    pool levels but also in a decrease in noncovalently bound heme

    (Figure 5). The lowering of heme levels in the chloroplasts after

    shift to light can be expected to correlate with an at least partial

    release of GluTR from feedback inhibition by heme (Vothknecht

    et al., 1996), resulting in a higher flux through the tetrapyrrole

    biosynthetic pathway and thus the accumulation of Proto,

    MgProto, and MgProtoMe (pathway 1 in Figure 11). The de-

    crease in heme levels in the presence of increasing Proto levels

    indicates a preferential channeling of Proto into chlorophyll bio-

    synthesis, which may be explained by the observed light activa-

    tion of CHL (Figure 6). (2) A negative regulator of tetrapyrrole

    biosynthesis that affects the Mg branch of tetrapyrrole biosyn-

    thesis is the Flu protein. Mutants of Arabidopsis that carry a

    lesion in the FLU gene are no longer able to downregulate

    d-amino levulinic acid synthesis and thus accumulate proto-

    chlorophyllide when grown in the dark (Meskauskiene et al.,

    2001). The Flu protein has subsequently been shown to interact

    with GluTR (Meskauskiene and Apel, 2002). A shift from dark to

    light enables catalysis of light-dependent protochlorophyllide

    reduction by protochlorophyllide oxidoreductase, and this, in the

    case of higher plants (Meskauskiene et al., 2001) and Chlamy-

    domonas (Pöpperl, 1996), has been shown to result in a reduc-

    tion in protochlorophyllide levels. In light, GluTR may be relieved

    from a negative control by the Flu protein, resulting in an

    enhanced synthesis of tetrapyrroles (pathway 2 in Figure 11).

    (3) There remains the possibility that early steps of tetrapyrrole

    biosynthesis are subject to direct light activation (e.g., via post-

    translational modification) (pathway 3 in Figure 11). Which of

    these routes or combination of routes results in the increased

    synthesis of tetrapyrroles observed upon illumination remains to

    be elucidated.

    In light of our model on the involvement of tetrapyrroles in

    chloroplast retrograde signaling in Chlamydomonas, we can now

    address the differences between green algae and higher plants.

    The recent identification of the Arabidopsis GUN1 gene and the

    analysis of mutants defective in this gene resulted in a model in

    which a chloroplast-localized protein with DNA/RNA binding

    activity (GUN1) generates a signal that coordinates nuclear gene

    expression in response to various alterations within the plastids

    (i.e., the accumulation of MgProto, inhibition of plastid gene ex-

    pression, and the redox state of the photosynthetic electron

    Figure 11. Model for the Activation of the HSP70A Promoter by MgProto and Heme.

    A strongly abbreviated pathway for the synthesis of heme and chlorophyll, starting with Glu, is indicated. (1) Feedback inhibition of GluTR by heme;

    (2) feedback regulation from the Mg-tetrapyrrole branch of the pathway mediated by the Flu protein (Goslings et al., 2004); (3) a direct activating effect

    of light on the first steps of tetrapyrrole biosynthesis. Details of the model are presented in the Discussion.

    562 The Plant Cell

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/20/3/552/6092166 by guest on 23 June 2021

  • transfer chain) (Koussevitzky et al., 2007). Our data obtained with

    the green alga Chlamydomonas reveal a number of differences:

    (1) The sequence motifs that meditate the response to elevated

    MgProto levels in higher plants and Chlamydomonas are not re-

    lated (Strand et al., 2003; von Gromoff et al., 2006; Koussevitzky

    et al., 2007). In Arabidopsis, the ABI4 transcription factor medi-

    ates the response to a signal from GUN1; in Chlamydomonas, the

    factor that binds PRE is still elusive. (2) Heme in Chlamydomonas

    has a regulatory function in gene regulation; in Arabidopsis, no

    evidence was observed for a role of heme similar to that of

    MgProto (i.e., in the repression of Lhcb) (Strand et al., 2003). (3)

    Within the Chlamydomonas HSP70A promoter, three different

    regions have been identified that mediate activation of the gene

    in response to tetrapyrroles, hydrogen peroxide, and singlet

    oxygen (Shao et al., 2007). This indicates that in Chlamydomonas

    different plastid signals appear not to converge in a common

    pathway. (4) A gene with homology to GUN1 (i.e., encoding a

    pentatricopeptide repeat protein with a mutS-related domain)

    was not detected in the Chlamydomonas genome sequence

    (http://genome.jgi-psf.org/Chlre3/Chlre3.home.html).

    How can the light induction of HSP70A, previously postulated

    to be mediated by MgProto, be explained in view of the reduced

    MgProto levels in the four CHL mutants? Here, we propose that,

    besides MgProto, heme may also mediate light induction. How-

    ever, this function of heme may be restricted to a special set of

    conditions (e.g., when Mg-tetrapyrrole levels are low and heme

    levels are elevated). The heme levels in the mutants are twofold

    to fivefold higher than in the wild type (Table 2), and they do not

    decrease upon shift of mutant cultures from dark to light (see

    Supplemental Figure 2 online). Since the mutants retain HSP70A

    activation in response to light, we postulate that heme in a light-

    dependent step is made available to a downstream signaling

    pathway in the cytosol/nucleus (see below) (Figure 11).

    The presence of a signaling pathway for heme, including a

    defined promoter element by which heme may modulate gene

    expression, is viewed as a strong indicator for a regulatory role of

    heme in Chlamydomonas. We envision that, besides a role in light

    induction of PRE-harboring genes, heme may have regulatory

    functions under physiological conditions when heme levels are

    elevated but MgProto is not available or strongly reduced. Such

    conditions in plants exist in non-green tissue but also have been

    documented for tobacco grown under a 12-h-light/12-h-dark

    cycle where pool levels of various tetrapyrroles were shown to

    change in opposite directions. Toward the end of the light period,

    pools of Proto, MgProto, and MgProtoMe were shown to drop

    approximately fivefold, while heme levels increased ;1.5-fold(Papenbrock et al., 1999).

    In Chlamydomonas, the chloroplast has been suggested to be

    the only site of heme synthesis (van Lis et al., 2005), and heme,

    also in the dark, will constantly be required to serve as prosthetic

    group of proteins in the cytosol and the mitochondria. Thus, a low

    level of heme must constantly exit from the chloroplast. Heme

    efflux from chloroplasts has indeed been observed (Thomas and

    Weinstein, 1990). We can envision that the Km of the heme

    binding factor(s) involved in signaling in the cytosol is significantly

    higher than that of reactions involved in the insertion of heme as a

    prosthetic group. The signaling factors thus only may become

    activated when the heme concentration in the cytosol/nucleus is

    raised (e.g., after an active, light-driven export of heme from the

    chloroplast). This hypothesis may be tested once heme binding

    factors have been identified.

    Information about the downstream signaling pathway was

    gained by experiments in which cultures were incubated with

    hemin for 12 h (Figure 10). This pathway was shown to share a

    genuine property of most signaling pathways (i.e., it is inactivated

    upon prolonged stimulation). In cells treated with hemin for 12 h,

    HSP70A light induction was essentially abolished (Figure 10).

    Since the addition of MgProto in the dark also did not elicit an

    induction of HSP70A, we may conclude that shared components

    of the downstream signaling pathway are blocked. The lack of

    HSP70A induction by either MgProto or hemin feeding suggests

    that either the receptor(s) for these tetrapyrroles or downstream

    components got inactivated. Indeed, the employment of the

    same cis-acting PRE element of the HSP70A promoter for the

    activation by both MgProto and heme (Figure 7) may point to a

    sharing of transcription factors by the two tetrapyrroles.

    These studies provide evidence for a regulatory role of heme in

    gene expression in Chlamydomonas. Whether this tetrapyrrole

    also is active in gene regulation in higher plants remains to be

    elucidated. In yeast and animal systems, a role of heme in

    regulation of gene expression is well established (reviewed in

    Mense and Zhang, 2006). Thus, in mammals, heme has recently

    been demonstrated to be an amplifier of the inflammatory

    response via specific activation of toll-like receptor 4 (Figueiredo

    et al., 2007). The details of the signaling pathways in Chlamydo-

    monas can now be addressed, and this may provide clues to a

    regulatory role of heme in higher plants.

    METHODS

    Algal Strains and Culture Conditions

    Chlamydomonas reinhardtii strain 325 (cw15, arg7-8, mtþ) was kindlyprovided by R. Matagne (University of Liège, Belgium). Wild-type strain

    CC-124 (mt�) and the cell wall–deficient mutant CW15 were obtainedfrom the Chlamydomonas Culture Collection at Duke University. Wild-

    type strain 4Aþ (mtþ) was provided by J.-D. Rochaix (University ofGeneva). The mutants were generated by M. Kobayashi in the 4Aþ strainby UV mutagenesis applying between 30 and 60 mJ cm�2. After muta-

    genesis, the cells were plated and incubated in the dark. Mutant clones

    were recognized by their altered pigmentation. If not otherwise noted,

    strains were grown photomixotrophically in Tris-acetate phosphate (TAP)

    medium (Harris, 1989) on a rotatory shaker at 238C under continuous

    irradiation with white light (;60 mE m�2 s�1) provided by fluorescenttubes (Osram L36W/25). Mutant stocks were maintained on TAP agar

    medium in the dark. Light induction and heat shock were performed

    according to Kropat et al. (1995) and von Gromoff et al. (1989), respec-

    tively. For Arg-requiring strains, the medium was supplemented with Arg

    (final concentration 100 mg/mL).

    Nuclear Transformation of Chlamydomonas

    Chlamydomonas nuclear transformation was performed using the glass

    bead method (Kindle, 1990), modified as described previously (Kropat

    et al., 2000). For the transformation of 108 cells, 100 ng of plasmid DNA or

    1 mg of BAC DNA were used. Immediately after vortexing with glass

    beads, cells were spread onto freshly prepared TAP-agar (1%) plates.

    Plates were incubated at 238C in the light (;60 mE m�2 s�1), and

    Heme Regulates Nuclear Genes 563

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/20/3/552/6092166 by guest on 23 June 2021

  • transformants were collected and analyzed after 1 week. For promoter

    activity tests, strain 325 was transformed as described by von Gromoff

    et al. (2006). The constructs (all containing the ARG7 gene) were linear-

    ized by restriction either upstream from HSP70A promoter sequences or

    downstream from the reporter gene.

    RNA Gel Blot Analysis

    Total RNA was isolated from 10 mL cultures grown to 2 to 4 3 106 cells per

    mL. The procedures employed for RNA extraction, separation on agarose

    gels, blotting, and hybridization were as described previously (von

    Gromoff et al., 1989) except that the nylon membranes used were

    Hybond-N (Amersham). The probes used for hybridization were as

    follows: HSP70A (von Gromoff et al., 1989), CHLH 1 (Chekounova

    et al., 2001), CHLI 1 (a SpeI-XhoI cDNA fragment of ;1.5 kb from ESTclone AV642320), CHLI 2 (a SpeI-XhoI fragment of ;1.8 kb from ESTclone AV623288), CHLD (a 680-bp cDNA fragment), and HEMA (a 560-bp

    cDNA fragment), kindly provided by S. Beale. The EST clones were

    obtained from the Japanese collection (Asamizu et al., 1999). Either

    CBLP, which encodes a Gb-like polypeptide (von Kampen et al., 1994), or

    a clone (Ba235) harboring the 18s and 25s rRNA genes (Marco and

    Rochaix, 1980) was used as loading controls. Probes were labeled with

    50 mCi [a-32P]dCTP (Amersham) according to the random priming pro-

    tocol (Feinberg and Vogelstein, 1983). Hybridizations were performed as

    described previously (Wegener and Beck, 1991). After hybridization, the

    membranes were washed twice in 23 SSC for 5 min at room temperature,

    once in 23 SSC, 1% SDS for 30 min at 658C, and once in 0.13 SSC, 1%

    SDS for 30 min at 658C. RNA gel blots were screened by exposing

    membranes to BAS-MP imaging plates (Fuji). They were evaluated by a

    phosphor imager (Bio-Rad Laboratories). For the quantitative determi-

    nation of changes in mRNA concentrations, all clones that exhibited a

    response were evaluated using the Quantity One 4, 5, 1 program from

    Bio-Rad Laboratories.

    RNA Dot Blot Analysis

    Ten micrograms of total RNA from each sample was applied to Hybond-N

    nylon membranes using a dot blot apparatus. The membranes were dried

    and RNA fixed to the membranes by baking for 2 h at 808C. For

    hybridization, washing of the membranes, and their quantitative evalua-

    tion, we used the same conditions/programs as for RNA gel blot analyses.

    Isolation of Genes for CHL

    BAC clones containing genes CHLH, CHLD, CHLI 1, and CHLI 2 were

    identified by screening the membrane-immobilized DNA of a Chlamydo-

    monas BAC library (Lefebvre and Silflow, 1999) obtained from Clemson

    University Genomics Institute (http://www.genome.clemson.edu/) using

    the same probes as described for RNA detection. Hybridization of the

    membranes was performed using the protocol provided by the Clemson

    University Genomics Institute. Positive BAC clones that contained the

    complete genes (17G11-CHLH 1, 26C5-CHLD, 35P13-CHLI 1, and 5J15-

    CHLI 2) were obtained from Clemson University Genomics Institute. DNA

    was isolated using the protocol of the BAC clone providers (http://

    www.genome.clemson.edu/protocols.shtml).

    Cloning and Sequencing of CHLD

    Genomic sequence data (http://genome.jgi-pst.org/Chlre3/Chlre3.

    home.html) provided information on restriction sites used for cloning of

    CHLD. An 8.2-kb fragment containing the entire CHLD gene was obtained

    by digestion of DNA of BAC clone 26C5 with NheI and HpaI. The resulting

    8.2-kb fragment after blunt ending was ligated into the EcoRV site of the

    pBluescript SKþ/� vector (Stratagene). The clone containing the correctinsert was verified by partial sequencing and restriction analysis. DNA of

    this clone was tested for its ability to complement mutants defective in

    chlorophyll synthesis upon transformation. For the identification of the

    mutations in two mutants that could be complemented by the CHLD gene,

    DNA was extracted from mutant strains chlD-1 and chlD-2 using the CTAB

    method (von Gromoff et al., 1989). The sequence of the CHLD mutant

    alleles was determined after amplification of segments that comprised the

    entire CHLD coding region, using six primer pairs. The upstream (A) and

    downstream (B) primers were as follows: 1A, 59-CACAACCAGGAC-

    AGCCTA-39; 1B, 59-ATCGAAGCCCTTGGAGTC-39; 2A, 59-TGTCTGAG-

    GAGGACTCCA-39; 2B, 59-GGGCTGGAACACAGTCTT-39; 3A, 59-AGG-

    GCAAGACTGTGTTCCA-39; 3B, 59-CGATGGTCACGTCCTTCA-39; 4A,

    59-TCATCCTGGCTCGCGAGTA-39; 4B, 59-CTTGTTCACACGCTCACG-39;

    5A, 59-TGTCTCGAGCGAGTTGCTG-39; 5B, 59-ATGCGCTGTACTCACT-

    GTCTG-39; 6A, 59-CTCATCTTCAGCGACGAC-39; 6B, 59-ACCTCAGTCA-

    CTCGGCA-39.

    The reactions were performed for 35 cycles of 1 min at 948C, 1 min at 50

    or 558C, and 2.5 min at 728C with Pfu DNA polymerase (Fermentas), which

    possesses proofreading activity. Resulting overlapping PCR products

    were subcloned into the pGEM-T vector system (Promega) and se-

    quenced. Sequencing was performed by the dideoxynucleotide chain

    termination method (Sanger et al., 1977) using the ALF DNA analyzing

    system (Amersham Biosciences Europe).

    Nucleic Acid Manipulations

    As the basic plasmid used for all promoter activity tests, we employed

    pCB412, which is pARG7.8cos (Purton and Rochaix, 1995) with a

    polylinker inserted into the unique NruI site. All constructs tested were

    located on this vector and have been described previously (von Gromoff

    et al., 2006).

    Sources of Porphyrins

    The MgProto used came either from Frontier Scientific or was prepared as

    described previously (Kropat et al., 1997). This latter MgProto was kindly

    provided by U. Oster. The concentration of MgProto was determined in

    ether by spectrophotometric measurements using a millimolar extinction

    coefficient of 308 at 419 nm (Falk, 1964). The concentration of hemin

    (Sigma-Aldrich) was determined in a solvent consisting of 66.5% ethanol,

    17% acetic acid, and 16.5% water (v/v/v) using a millimolar extinction

    coefficient of 144 at 398 nm (Thomas and Weinstein, 1990). Other

    metalated porphyrins were from Frontier Scientific. Their concentrations

    were determined as specified by the supplier. For tests in vivo, tetrapyr-

    roles were dissolved in dimethyl sulfoxide and used at the final concen-

    tration indicated. The final dimethyl sulfoxide concentration in the culture

    medium in all cases was 0.25%.

    Pigment Analyses

    Porphyrin steady state levels were determined in cells grown in liquid.

    Cells were harvested and stored at�808C prior to rupture by sonication inan ice bath. Porphyrins were extracted in three steps with (1) methanol, (2)

    K-phosphate buffer, and (3) a mixture of acetone: methanol: 0.1 N NH4OH

    (10:9:1, v/v) (Alawady and Grimm, 2005). Aliquots of the supernatant were

    separated by HPLC (model 1100; Agilent) on an RP 18 column (Novapak

    C18, 4 mm particle size, 3.9 3 150 mm; Waters Chromatography) at a flow

    rate of 1 mL/min in a methanol/0.1 M ammonium acetate, pH 5.2, gradient

    and eluted with a linear gradient of solvent B (90% methanol and 0.1 M

    ammonium acetate, pH 5.2) and solvent A (10% methanol and 0.1 M

    ammonium acetate, pH 5.2). The eluted samples were monitored by a

    564 The Plant Cell

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/20/3/552/6092166 by guest on 23 June 2021

  • fluorescence detector at an excitation wavelength of 405 nm and emis-

    sion wavelength of 620 nm for Proto and an excitation wavelength of 420

    nm and emission wavelength of 595 nm for Mg-porphyrins. The porphy-

    rins and metalloporphyrins were identified and quantified by comparison

    with authentic standards purchased from Frontier Scientific. The content

    of noncovalently bound heme was determined after removal of chloro-

    phyll from Chlamydomonas cells by intensive washing with ice-cold

    alkaline acetone containing 0.1 N NH4OH (9:1; v/v) (Weinstein and Beale,

    1984). Heme was then extracted with acidic acetone containing 5% HCl,

    transferred to diethyl ether, concentrated, and washed on a DEAE-

    Sepharose column. Absorbance was monitored at 398 nm using a milli-

    molar extinction coefficient of 144. The concentration of heme standard

    solutions was determined spectrophotometrically (Weinstein and Beale,

    1983).

    Determination of CHL Activity

    For determination of CHL activity, Chlamydomonas cells were grown in

    500 mL of liquid TAP medium to a density of 3 to 5 3 106 cells per mL,

    harvested by centrifugation, and resuspended in homogenization buffer

    (0.5 M sorbitol, 0.1 M Tris-HCl, 1 mM DTT, and 0.1% BSA, pH 7.5). The

    cells were disrupted by sonication in an ice bath. The cell extracts were

    centrifuged at 15,000g for 30 min and the supernatants were used for

    enzyme assays. CHL was assayed as described by Walker and Weinstein

    (1991). The assay mixture contained 40 mM MgCl2 and 20 mM ATP. The

    reaction, performed at 288C in individual opaque tubes, was started by

    adding 100 mM Proto and stopped after 5, 10, 15, 30, and 60 min by

    freezing the tubes in liquid nitrogen. Samples from each enzyme assay

    were extracted and the products quantified according to methods de-

    scribed above.

    Accession Numbers

    Sequence data from this article can be found in the GenBank/EMBL data

    libraries under accession numbers AJ307055 (CHLH) and AF343974

    (CHLI1). The other genes may be found as gene models in the Chlamy-

    domonas genome sequence (http://genome.jgi-psf.org/Chlre3/Chlre3.

    home.html) (CHLI2, gene model C_110312; CHLD, gene model C_60238).

    Supplemental Data

    The following materials are available in the online version of this article.

    Supplemental Figure 1. Test of the Effect of Inhibitors of Cytosolic

    and Organellar Protein Synthesis on the Light Activation of Mg-

    Chelatase Genes.

    Supplemental Figure 2. Assay of Tetrapyrrole Pool Levels in Mg-

    Chelatase Mutants after a Shift from Dark to Light.

    Supplemental Figure 3. Assay of Tetrapyrrole Pool Levels in Wild-

    Type Cells That, before a Dark-to-Light Shift, Were Incubated with

    Hemin in the Dark for 12 h.

    ACKNOWLEDGMENTS

    We thank Krishna Niyogi and Marilyn Kobayashi (University of California,

    Berkeley, CA), in whose laboratory the mutants were isolated during a

    sabbatical stay of C.F.B. in Berkeley. We also thank Ulrike Oster for

    providing us with MgProto and Wolfhart Rüdiger (both University of

    Munich) for helpful suggestions. Michael Schroda’s (University of Frei-

    burg) valuable suggestions and his critical reading of the manuscript are

    greatly appreciated. This work was supported by a grant from the

    Deutsche Forschungsgemeinschaft to C.F.B.

    Received July 31, 2007; revised February 15, 2008; accepted February

    29, 2008; published March 25, 2008.

    REFERENCES

    Abdallah, F., Salamini, F., and Leister, D. (2000). A prediction of the

    size and evolutionary origin of the proteome of chloroplasts of

    Arabidopsis. Trends Plant Sci. 5: 141–142.

    Alawady, A.E., and Grimm, B. (2005). Tobacco Mg protoporphyrin IX

    methyl-transferase is involved in inverse activation of Mg porphyrin

    and protoheme synthesis. Plant J. 41: 282–290.

    Ankele, E., Kindgren, P., Pesquet, E., and Strand, Å. (2007). In vivo

    visualization of Mg-ProtoporphyrinIX, a coordinator of photosynthetic

    gene expression in the nucleus and the chloroplast. Plant Cell 19:

    1964–1979.

    Asamizu, E., Nakamura, Y., Sato, S., Fukuzawa, H., and Tabata, S.

    (1999). A large scale structural analysis of cDNAs in a unicellular green

    alga, Chlamy-domonas reinhardtii. I. Generation of 3433 non-redundant

    expressed sequence tags. DNA Res. 6: 369–373.

    Balmer, Y., Koller, A., del Val, G., Manieri, W., Schürmann, P., and

    Buchanan, B.B. (2003). Proteomics gives insight into the regulatory

    function of chloroplast thioredoxins. Proc. Natl. Acad. Sci. USA 100:

    370–375.

    Beale, S.I. (1999). Enzymes of chlorophyll biosynthesis. Photosynth.

    Res. 60: 43–73.

    Beck, C.F. (2005). Signaling pathway from the chloroplast to the

    nucleus. Planta 222: 743–756.

    Beck, C.F., and Grimm, B. (2006). Involvement of tetrapyrroles in

    cellular regulation. In Biochemistry, Biophysics and Biological Func-

    tions of Chlorophylls, B. Grimm, R.J. Porra, W. Rüdiger, and H.

    Scheer, eds (Dordrecht, The Netherlands: Springer), pp. 223–233.

    Chekounova, E., Voronetskaya, V., Papenbrock, J., Grimm, B., and

    Beck, C.F. (2001). Characterization of Chlamydomonas mutants

    defective in the H-subunit of Mg-chelatase. Mol. Genet. Genomics

    266: 363–373.

    Falk, J.E. (1964). Porphyrins and Metalloporphyrins: Their General,

    Physical and Coordination Chemistry, and Laboratory Methods.

    (Amsterdam, The Netherlands: Elsevier Publishing).

    Feinberg, A.P., and Vogelstein, B. (1983). A technique for radiolabel-

    ling DNA restriction endonuclease fragments to high activity. Anal.

    Biochem. 132: 2–13.

    Figueiredo, R.T., Fernandez, P.L., Mourao-Sa, D.S., Porto, B.N.,

    Dutra, F.F., Alves, L.S., Oliveira, M.F., Oliveira, P.L., Graça-Souza,

    A.V., and Bozza, M.T. (2007). Characterization of heme as activator

    of toll-like receptor 4. J. Biol. Chem. 282: 20221–20229.

    Goldschmidt-Clermont, M. (1998). Coordination of nuclear and chlo-

    roplast gene expression in plant cells. Int. Rev. Cytol. 177: 115–180.

    Goslings, D., Meskauskiene, R., Kim, C., Lee, K.P., Nater, M., and

    Apel, K. (2004). Concurrent interactions of heme and FLU with Glu

    tRNA reductase (HEMA1), the target of metabolic feedback inhibition

    of tetrapyrrole biosynthesis, in dark- and light-grown Arabidopsis

    plants. Plant J. 40: 957–967.

    Gray, J.C., Sullivan, J.A., Wang, J.H., Jerome, C.A., and MacLean, D.

    (2003). Coordination of plastid and nuclear gene expression. Philos.

    Trans. R. Soc. Lond. B Biol. Sci. 358: 135–144.

    Harris, E.H. (1989). The Chlamydomonas Source Book: A Comprehen-

    sive Guide to Biology and Laboratory Use. (San Diego, CA: Academic

    Press).

    Ikegami, A., Yoshimura, N., Motohashi, K., Takahashi, S., Romano,

    P.G.N., Hisabori, T., Takamiya, K., and Masuda, T. (2007). The

    CHLI1 subunit of Arabidopsis thaliana magnesium chelatase is a target

    protein of the chloroplast thioredoxin. J. Biol. Chem. 282: 19282–19291.

    Heme Regulates Nuclear Genes 565

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/20/3/552/6092166 by guest on 23 June 2021

  • Kindle, K.L. (1990). High-frequency nuclear transformation of Chlamy-

    domonas reinhardtii. Proc. Natl. Acad. Sci. USA 87: 1228–1232.

    Koussevitzky, S., Nott, A., Mockler, T.C., Hong, F., Sachetto-

    Martins, G., Surpin, M., Lim, I.J., Mittler, R., and Chory, J. (2007).

    Signals from chloroplasts converge to regulate nuclear gene expres-

    sion. Science 316: 715–719.

    Kropat, J., Oster, U., Rüdiger, W., and Beck, C.F. (1997). Chlorophyll

    precursors are signals of chloroplast origin involved in light induction

    of nuclear heat-shock genes. Proc. Natl. Acad. Sci. USA 94: 14168–

    14172.

    Kropat, J., Oster, U., Rüdiger, W., and Beck, C.F. (2000). Chloroplast

    signaling in the light induction of nuclear HSP70 genes requires the

    accumulation of chlorophyll precursors and their accessibility to

    cytoplasm/nucleus. Plant J. 24: 523–531.

    Kropat, J., von Gromoff, E.D., Müller, F.W., and Beck, C.F. (1995).

    Heat shock and light activation of a Chlamydomonas HSP70 gene are

    mediated by independent regulatory pathways. Mol. Gen. Genet. 248:

    727–734.

    La Rocca, N., Rascio, N., Oster, U., and Rüdiger, W. (2001). Amitrole

    treatment of etiolated barley seedlings leads to deregulation of

    tetrapyrrole synthesis and to reduced expression of Lhc and RbcS

    genes. Planta 213: 101–108.

    Lefebvre, P.A., and Silflow, C.D. (1999). Chlamydomonas: The cell and

    its genome. Genetics 151: 9–14.

    Leister, D. (2003). Chloroplast research in the genomic age. Trends

    Genet. 19: 47–56.

    Leon, P., Arroyo, A., and Mackenzie, S. (1998). Nuclear control of

    plastid and mitochondrial development in higher plants. Annu. Rev.

    Plant Physiol. Plant Mol. Biol. 49: 453–480.

    Lohr, M., Im, C.-S., and Grossman, A.R. (2005). Genome-based

    examination of chlorophyll and carotenoid biosynthesis in Chlamydo-

    monas reinhardtii. Plant Physiol. 138: 490–515.

    Marco, Y., and Rochaix, J.D. (1980). Organization of the nuclear

    ribosomal DNA of Chlamydomonas reinhardtii. Mol. Gen. Genet. 177:

    715–723.

    Martin, W., Rujan, T., Richly, E., Hansen, A., Cornelsen, S., Lins, T.,

    Leister, D., Stoebe, B., Hasegawa, M., and Penny, D. (2002).

    Evolutionary analysis of Arabidopsis, cyanobacterial, and chloroplast

    genomes reveals plastid phylogeny and thousands of cyanobacterial

    genes in the nucleus. Proc. Natl. Acad. Sci. USA 99: 12246–12251.

    Maul, J.E., Lilly, J.W., Cui, L., dePhamphilis, C.W., Miller, W., Harris,

    E.H., and Stern, D.B. (2002). The Chlamydomonas reinhardtii plastid

    chromosome: islands of genes in a sea of repeats. Plant Cell 14:

    2659–2679.

    Mense, S.M., and Zhang, L. (2006). Heme: A versatile signaling

    molecule controlling the activities of diverse regulators ranging from

    transcription factors to MAP kinases. Cell Res. 16: 681–692.

    Meskauskiene, R., and Apel, K. (2002). Interaction of FLU, a negative

    regulator of tetrapyrrole biosynthesis, with the glutamyl-tRNA reduc-

    tase requires the tetratricopeptide repeat domain of FLU. FEBS Lett.

    532: 27–30.

    Meskauskiene, R., Nater, M., Goslings, D., Kessler, D., op den

    Camp, R., and Apel, K. (2001). Flu: A negative regulator of chlorophyll

    biosynthesis in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 98:

    12826–12831.

    Mochizuki, N., Brusslan, J.A., Larkin, R., Nagatani, A., and Chory, J.

    (2001). Arabidopsis genomes uncoupled 5 (GUN 5) mutant reveals the

    involvement of Mg-chelatase H subunit in plastid-to-nucleus signal

    transduction. Proc. Natl. Acad. Sci. USA 98: 2053–2058.

    Nott, A., Jung, H.-S., Koussevitzky, S., and Chory, J. (2006). Plastid-

    to-nucleus retrograde signaling. Annu. Rev. Plant Biol. 57: 739–759.

    Oster, U., Brunner, H., and Rüdiger, W. (1996). The greening process

    in cress seedlings. V. Possible interference of chlorophyll precursors,

    accumulated after thujaplicin treatment, with light-regulated expres-

    sion of Lhc genes. J. Photochem. Photobiol. B 36: 255–261.

    Papenbrock, J., and Grimm, B. (2001). Regulator network of tetrapyr-

    role biosynthesis – Studies for intracellular signalling involved in

    metabolic and developmental control of plastids. Planta 213:

    667–681.

    Papenbrock, J., Mock, H.-P., Kruse, E., and Grimm, B. (1999).

    Expression studies in tetrapyrrole biosynthesis: inverse maxima of

    magnesium chelatase and ferrochelatase activity during cyclic pho-

    toperiods. Planta 208: 264–273.

    Papenbrock, J., Mock, H.-P., Tanaka, R., Kruse, E., and Grimm, B.

    (2000a). Role of magnesium chelatase activity in the early steps of the

    tetrapyrrole biosyn-thetic pathway. Plant Physiol. 122: 1161–1169.

    Papenbrock, J., Pfündel, E., Mock, H.-P., and Grimm, B. (2000b).

    Decreased and increased expression of the subunit CHL I diminishes

    Mg-chelatase activity and reduces chlorophyll synthesis in transgenic

    tobacco plants. Plant J. 22: 155–164.

    Pfannschmidt, T. (2003). Chloroplast redox signals: How photosynthe-

    sis controls its own genes. Trends Plant Sci. 1: 33–41.

    Pontier, D., Albrieux, C., Joyard, J., Lagrange, T., and Block, M.A.

    (2007). Knock-out of the magnesium protoporphyrin IX methyltrans-

    ferase gene in Arabidopsis. Effects on chloroplast development and

    on chloroplast-to-nucleus signaling. J. Biol. Chem. 282: 2297–2304.

    Pöpperl, G. (1996). Die Analytik von Chlorophyll-Vorstufen: Anwendung

    auf die Untersuchung der Lichtregulation der Biosynthese von Chlo-

    rophyll. PhD dissertation (Munich, Germany: University of Munich).

    Pöpperl, G., Oster, U., and Rüdiger, W. (1998). Light-dependent

    increase in chlorophyll precursors during the day-night cycle in

    tobacco and barley seedlings. J. Plant Physiol. 153: 40–45.

    Purton, S., and Rochaix, J.-D. (1995). Characterization of the ARG7

    gene of Chlamydomonas reinhardtii and its application to nuclear

    transformation. Eur. J. Phycol. 30: 141–148.

    Rodermel, S. (2001). Pathway of plastid-to-nucleus signalling. Trends

    Plant Sci. 6: 471–478.

    Sanger, F., Nicklen, S., and Coulson, A.R. (1977). DNA sequence with

    chain-terminating inhibitors. Proc. Natl. Acad. Sci. USA 74: 5463–

    5467.

    Shao, N., Krieger-Liszkay, A., Schroda, M., and Beck, C.F. (2007). A

    reporter system for the individual detection of hydrogen peroxide and

    singlet oxygen: its use for the assay of reactive oxygen species

    produced in vivo. Plant J. 50: 475–487.

    Silflow, C.D. (1998). Organization of the nuclear genome. In Molecular

    Biology of Chlamydomonas: Chloroplasts and Mitochondria, J.-D.

    Rochaix, M. Goldschmidt-Clermont, and S. Merchant, eds (Dor-

    drecht, The Netherlands: Kluwer), pp. 25–40.

    Strand, A. (2004). Plastid-to-nucleus signalling. Curr. Opin. Plant Biol. 7:

    621–625.

    Strand, A., Asami, T., Alonso, J., Ecker, J.R., and Chory, J. (2003).

    Chloroplast to nucleus communication triggered by accumulation of

    Mg-protoporphyrinIX. Nature 421: 79–83.

    Susek, R.E., Ausubel, F.M., and Chory, J. (1993). Signal transduction

    mutants of Arabidopsis uncouple nuclear CAB and RBCS gene

    expression from chloroplast development. Cell 74: 787–799.

    Thomas, J., and Weinstein, J.D. (1990). Measurement of heme efflux

    and heme content in isolated developing chloroplasts. Plant Physiol.

    94: 1414–1423.

    van Lis, R., Atteia, A., Nogaj, L.A., and Beale, S.I. (2005). Subcellular

    localization and light-regulated expression of protoporphyrinogen IX

    oxidase and ferroche-latase in Chlamydomonas reinhardtii. Plant

    Physiol. 139: 1946–1958.

    Vasileuskaya, Z., Oster, U., and Beck, C.F. (2004). Involvement of

    tetrapyrroles in inter-organellar signaling in plants and algae. Photo-

    synth. Res. 82: 289–299.

    566 The Plant Cell

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/20/3/552/6092166 by guest on 23 June 2021

  • Vasileuskaya, Z., Oster, U., and Beck, C.F. (2005). Identification of

    Mg-protoporphyrinIX and heme as plastid/mitochondrial factors that

    control HEMA expression in Chlamydomonas. Eukaryot. Cell 4: 1620–

    1628.

    von Gromoff, E.D., Schroda, M., Oster, U., and Beck, C.F. (2006).

    Identification of a plastid response element that acts as an enhancer

    within the Chlamydomonas HSP 70A promoter. Nucleic Acids Res.

    34: 4767–4779.

    von Gromoff, E.D., Treier, U., and Beck, C.F. (1989). Three light-

    inducible heat shock genes of Chlamydomonas reinhardtii. Mol. Cell.

    Biol. 9: 3911–3918.

    von Kampen, J., Nieländer, U., and Wettern, M. (1994). Stress-

    dependent transcription of a gene encoding a Gb-like polypeptide

    from Chlamydomonas reinhardtii. J. Plant Physiol. 143: 756–758.

    Vothknecht, U.C., Kannangara, C.G., and von Wettstein, D. (1996).

    Expression of catalytically active barley glutamyl tRNAGlu reductase in

    Escherichia coli as a fusion protein with glutathione S-transferase.

    Proc. Natl. Acad. Sci. USA 93: 9287–9291.

    Walker, C.J., and Weinstein, J.D. (1991). Further characterization of

    the magnesium chelatase in isolated developing cucumber chloro-

    plasts: substrate specificity, regulation, intactness, and ATP require-

    ments. Plant Physiol. 95: 1189–1196.

    Walker, C.J., and Willows, R.D. (1997). Mechanism and regulation of

    Mg-chelatase. Biochem. J. 327: 321–333.

    Wang, W.-Y., Wang, W.I., Boynton, J.E., and Gillham, N.W. (1974).

    Genetic control of chlorophyll biosynthesis in Chlamydomonas. J. Cell

    Biol. 63: 806–823.

    Wegener, D., and Beck, C.F. (1991). Identification of novel genes

    specifically expressed in Chlamydomonas reinhardtii zygotes. Plant

    Mol. Biol. 16: 937–946.

    Weinstein, J.D., and Beale, S.I. (1983). Separate physiological roles

    and subcellular compartments for two tetrapyrrole biosynthetic path-

    ways in Euglena gracilis. J. Biol. Chem. 258: 6799–6807.

    Weinstein, J.D., and Beale, S.I. (1984). Biosynthesis of protoheme and

    heme as precursors solely from glutamate in the unicellular red alga

    Cyanidium caldarium. Plant Physiol. 74: 146–151.

    Willows, R.D., Gibson, L.C., Kanangara, C.G., Hunter, C.N., and von

    Wettstein, D. (1996). Three separate proteins constitute the magnesium

    chelatase of Rhodobacter sphaeroides. Eur. J. Biochem. 235: 438–443.

    Heme Regulates Nuclear Genes 567

    Dow

    nloaded from https://academ

    ic.oup.com/plcell/article/20/3/552/6092166 by guest on 23 June 2021