Skip to main content
Molecular Plant logoLink to Molecular Plant
. 2014 Mar 3;7(7):1105–1120. doi: 10.1093/mp/ssu022

The Translational Apparatus of Plastids and Its Role in Plant Development

Nadine Tiller 1, Ralph Bock 1,1
PMCID: PMC4086613  PMID: 24589494

SUMMARY

Systematic reverse genetic approaches in the nuclear and chloroplast genomes have greatly increased our knowledge about the structure, function, and biogenesis of chloroplast ribosomes, and about the molecular mechanisms of plastid protein biosynthesis. They also provided new insights into the regulation of plant development by the activity of plastid gene expression. We review our current knowledge about the translational apparatus of plastids and the impact of plastid translation on plant anatomy and plant morphology.

Key words: plastid, translation, ribosome, ribosomal protein, evolution, plastid transformation, retrograde signaling, leaf development, palisade cell.

Abstract

Chloroplasts (plastids) possess a genome and their own machinery to express it. Translation in plastids occurs on bacterial-type 70S ribosomes utilizing a set of tRNAs that is entirely encoded in the plastid genome. In recent years, the components of the chloroplast translational apparatus have been intensely studied by proteomic approaches and by reverse genetics in the model systems tobacco (plastid-encoded components) and Arabidopsis (nucleus-encoded components). This work has provided important new insights into the structure, function, and biogenesis of chloroplast ribosomes, and also has shed fresh light on the molecular mechanisms of the translation process in plastids. In addition, mutants affected in plastid translation have yielded strong genetic evidence for chloroplast genes and gene products influencing plant development at various levels, presumably via retrograde signaling pathway(s). In this review, we describe recent progress with the functional analysis of components of the chloroplast translational machinery and discuss the currently available evidence that supports a significant impact of plastid translational activity on plant anatomy and morphology.

INTRODUCTION

Chloroplasts (plastids) are DNA-containing cell organelles that are bounded by a double membrane. They evolved from a photosynthetically active cyanobacterium by an endosymbiotic event that took place more than a billion years ago (Gray, 1993). During evolution, the vast majority of the genetic information of the initially genetically autonomous endosymbiont was relocated to the nucleus of the host cell (Timmis et al., 2004; Bock and Timmis, 2008). However, a specific set of genes was retained in the plastid compartment and is expressed by a dedicated gene expression machinery. In most photosynthetically active vascular plants, the circularly mapping plastid genome (plastome) encodes approximately 120 genes that are densely arranged on a single chromosome of 120–220kb (reviewed in Wakasugi et al., 2001; Bock, 2007). The genes can be grouped into three major classes: genes encoding parts of the photosynthetic apparatus, genes encoding components of the genetic system, and genes encoding other functions (Shimada and Sugiura, 1991). Despite more than a billion years of separate evolution, the gene expression machinery of plastids still shares substantial similarities with that of its cyanobacterial ancestor, but also has acquired a number of novel organelle-specific features, components, and regulatory mechanisms. It seems reasonable to assume that many of these evolutionary novelties serve the purpose of optimizing the coordination of the expression of the chloroplast genome with that of the two other genetic compartments in the cell: the nucleus and the mitochondrion. In addition, the evolution of multicellularity and, in particular, the evolution of complex body plans in vascular plants posed new regulatory challenges related to the tissue-specific and developmental stage-specific differentiation of plastids into specialized plastid types (e.g. proplastids of meristematic tissues, chromoplasts of fruits and flowers, amyloplasts of roots and tubers).

EXPRESSION OF THE PLASTID GENOME

Plastid-encoded genes are transcribed by two types of RNA polymerases, both of which are necessary for the biogenesis of photosynthetically active chloroplasts (Allison et al., 1996; Swiatecka-Hagenbruch et al., 2008). The plastome itself encodes a bacterial-type multi-subunit RNA polymerase (plastid-encoded RNA polymerase, PEP) that requires nucleus-encoded sigma factors to facilitate promoter recognition. Additionally, a single subunit nucleus-encoded RNA polymerase (NEP) related to the RNA-synthesizing enzymes of T-type bacteriophages and mitochondria is present in plastids (Hedtke et al., 2002). In plastids of vascular plants, two such NEP enzymes are present (dubbed RPOTp and RPOTmp; Liere and Börner, 2007).

Primary transcripts produced by both polymerases are usually polycistronic and undergo extensive posttranscriptional processing steps, including intron removal by splicing, processing of primary polycistronic RNA molecules into mature monocistronic or oligocistronic mRNAs, trimming of the 5’ and 3’ ends, and RNA editing—a process by which the identity of cytidine residues at highly specific sites is altered to uridine (reviewed, e.g., in Bock, 2000; Stern et al., 2010; Barkan, 2011; Stoppel and Meurer, 2012). Members of the RNA-binding pentatricopeptide repeat protein (PPR) family have been shown to play crucial roles in many of these posttranscriptional processes (Schmitz-Linneweber and Small, 2008; Barkan, 2011).

Although significant transcriptional regulation occurs in plastids (Mullet and Klein, 1987; Allison and Maliga, 1995; Liere and Börner, 2007), it is generally believed that, in the course of evolution, the regulation of plastid gene expression was shifted from a predominantly transcriptional to a predominantly posttranscriptional control level (Eberhard et al., 2002). Posttranscriptional control is exerted at the level of RNA stability and especially at the level of translation (Mullet and Klein, 1987; Staub and Maliga, 1993; Eberhard et al., 2002; Kahlau and Bock, 2008). Plastid mRNAs are stabilized by sequence elements located in their 5’ and 3’ untranslated regions as well as by RNA secondary structures and RNA-binding PPR proteins (Stern et al., 2010; Barkan, 2011). Degradation is mediated by a complex interplay of endo- and exoribonucleases with removal of protective terminal stem-loop-type RNA secondary structures and 3’ polyadenylation being decisive initiating events (Stern et al., 2010; Stoppel and Meurer, 2012).

Translation of the genetic information in plastids is performed by ribosomes that are very similar to bacterial 70S ribosomes (Yamaguchi et al., 2000; Yamaguchi and Subramanian, 2000; Beligni et al., 2004b; Manuell et al., 2007). Almost all components of the plastid translational apparatus have close homologs in cyanobacteria, suggesting that the basic functions have been conserved throughout evolution. Translation initiation starts with the formation of a pre-initiation complex composed of the 30S ribosomal subunit and the initiator transfer RNA (tRNA) that selects the translation initiation site in the mRNA. AUG, and in rare cases the alternative triplets GUG or UUG, serve as initiation codons in plastid mRNAs (Sugiura et al., 1998). The precise molecular mechanism by which the plastid pre-initiation complex recognizes the initiator codon is not fully understood. In bacteria, the most prominent mechanism of translation initiation is based on binding of the anti-Shine–Dalgarno sequence (residing at the 3’ end of the 16S rRNA in the 30S ribosomal subunit) to a purine-rich sequence motif 7±2 nucleotides upstream of the initiator codon in the mRNA. This motif has been dubbed the Shine–Dalgarno sequence (or ribosome-binding site, RBS). While some plastid genes clearly possess bacterial-type Shine–Dalgarno sequences and are likely to utilize them for translation initiation (Ruf and Kössel, 1988; Hirose et al., 1998; Drechsel and Bock, 2010), many genes lack properly positioned Shine–Dalgarno sequences. A recent study showed that, in both chloroplasts and bacteria, translation initiation regions lacking Shine–Dalgarno motifs are less structured than those harboring a canonical Shine–Dalgarno sequence (Scharff et al., 2011). Absence of secondary structure formation around the initiation codon was proposed to facilitate start site recognition and binding of the 30S ribosomal subunit.

Following selection of the start codon, recruitment of the 50S subunit converts the pre-initiation complex into an active initiation complex, which then can enter the elongation phase of protein synthesis. Similarly to bacteria, the process is likely to be assisted by translation initiation and elongation factors. Chloroplast homologs of bacterial initiation and elongation factors have been identified and, in part, characterized (Lin et al., 1996; Beligni et al., 2004a; Albrecht et al., 2006; Miura et al., 2007; Ruppel and Hangarter, 2007; Shen et al., 2013). During translation elongation, the ribosome slides along the coding sequence of the mRNA incorporating the amino acids corresponding to each codon into the growing polypeptide chain with the help of the translation elongation factors EF-Tu, EF-G, and EF-Ts. Plastids use the standard genetic code, but it must be borne in mind that single nucleotide deviations between DNA sequence and mRNA sequence can be caused by RNA editing (Bock, 2000). Translation is terminated when the ribosome reaches one of the three standard termination codons (UAA, UAG, or UGA). The termination process requires the assistance of ribosome release factors (Meurer et al., 2002; Motohashi et al., 2007). To prepare the ribosome for a new round of protein synthesis, the mRNA–ribosome complex is subsequently disassembled by the ribosome-recycling factor (RRF; Wang et al., 2010). RRF dissociates the 70S ribosome into its 30S and 50S subunits, thereby releasing the mRNA and making the free 30S subunit available for de novo translation initiation.

The translation rate of individual mRNAs is mainly regulated at the level of translation initiation. Suggested control mechanisms include redox regulation (that may couple translation to photosynthetic electron transfer) and the so-called control by epistasy of synthesis (CES), an autoregulation mechanism that couples translation to protein complex assembly (Choquet et al., 1998; Peled-Zehavi and Danon, 2007).

The extensive regulation of plastid gene expression at the translational level calls for experimental methods that are suitable to measure chloroplast translational activity. Pulse labeling of newly synthesized chloroplast proteins using radioactively labeled amino acids (35S-labeled methionine and/or cysteine) represents one of the classical techniques that has been used in microorganisms and, to a much lesser extent, also in plants (e.g. Meurer et al., 1998b). However, with this method, only highly abundant proteins like Rubisco (RbcL), the large subunits of photosystem I (PsaA, PsaB), and photosystem II (D1, D2, CP43, CP47) can be readily detected. Also, in higher plants, it is difficult if not impossible to ensure the fast and homogeneous uptake of radiolabeled amino acids by multicellular tissues or even intact plants, making it hard to draw reliable quantitative conclusions from such experiments. Another frequently used technique to analyze plastid translational activity relies on the isolation of polysomes (Barkan, 1998). Polysomes are complexes of mRNAs and actively translating ribosomes. The number of ribosomes associated with a particular mRNA can serve as a proxy measure of the efficiency of its translation. Highly translated mRNAs are loaded with many ribosomes and can be separated from poorly translated and free mRNAs by analytical sucrose density centrifugation. The distribution of individual mRNAs across the gradient fractions is then assessed by Northern blot analysis (Barkan, 1988). By combining polysome isolation with microarray hybridization techniques, the method can be employed as a tool for the genome-wide analysis of translational regulation (also referred to as translatomics; Kahlau and Bock, 2008). This analytical platform has been used in a number of studies to examine the translational regulation of plastid gene expression in response to developmental cues and genetic perturbations (Kahlau and Bock, 2008; Valkov et al., 2009; Walter et al., 2010). Recently, a novel technique combining ribosomal footprinting with an oligonucleotide tiling array of the plastid ORFeome has been developed and successfully applied to determine the abundance and translational status of all chloroplast mRNAs in translation mutants of maize (Zoschke et al., 2013). Future research on chloroplast translation should benefit greatly from combining ribosomal footprinting with next-generation sequencing techniques (RNAseq) for translatomic analyses (Ingolia et al., 2009, 2012).

In view of the importance of translational regulation in the plastid (which can easily override even large changes in mRNA abundance; Eberhard et al., 2002), interfering with translation is the most appropriate reverse genetic method if down-regulation of a chloroplast gene or open reading frame is to be attempted. Knockdown of plastid genes by stable transformation of the plastid genome is a suitable strategy to study the function of essential genes whose knockout is lethal and therefore does not produce analyzable mutants (Drescher et al., 2000; Shikanai et al., 2001; Kode et al., 2005). It also provides a very valuable tool for the in-depth functional analysis of non-essential genes in that it can produce a spectrum of mutant phenotypes (Rott et al., 2011), which often are more informative than a complete gene knockout. So far, two strategies have been proven to be suitable to down-regulate the efficiency of translation in plastids: (1) the change of the standard translation initiation codon AUG to less efficient start codons like AUU, GUG, or UUG (Majeran et al., 2000; Hirose and Sugiura, 2004; Rott et al., 2011), and (2) the introduction of point mutations into the Shine–Dalgarno sequence (Hirose and Sugiura, 2004). Unfortunately, the recognition efficiency of non-AUG start codons is sequence-context-dependent (Boeck and Kolakofsky, 1994) and the same is likely to hold true for the interaction between the Shine–Dalgarno sequence and the anti-Shine–Dalgarno sequence in the 16S rRNA. Therefore, it is currently not possible to predict the efficacy of point mutations that are introduced into the translational start codon or the Shine–Dalgarno sequence of a plastid gene of interest. Consequently, different mutations need to be tested and, in view of the laborious and time-consuming procedures involved in stable transformation of the plastid genome, this represents a serious limitation.

ESSENTIAL AND NON-ESSENTIAL COMPONENTS OF THE PLASTID TRANSLATION MACHINERY

Ribosomal RNAs

The 70S-type chloroplast ribosome (sometimes also referred to as the ‘chlororibosome’) consists of two multi-component subunits: the large (50S) and the small (30S) ribosomal subunit. Both subunits are ribonucleoprotein complexes comprising one or more ribosomal RNA species (rRNAs) and many proteins. All ribosomal RNAs fold into highly complex three-dimensional structures and each rRNA species is associated with a distinct set of ribosomal proteins. The RNA components of the plastid ribosome are strikingly similar to their bacterial counterparts—a finding that was one of the first compelling molecular evidences for the endosymbiotic origin of plastids (Schwarz and Kössel, 1979, 1980). The 30S particle harbors a single rRNA molecule, the 16S rRNA, whereas the 50S particle contains three rRNAs: the 23S, 5S, and 4.5S rRNAs. A 4.5S rRNA does not exist in bacteria, but the plastid 4.5S rRNA is homologous to the 3’-terminal part of the bacterial 23S rRNA (Edwards et al., 1981), suggesting that it evolved by fragmentation of the 23S rRNA. Whether or not this split has any functional significance is currently not known.

All plastid ribosomal RNAs are encoded in the chloroplast genome, transcribed from a single operon (rrn operon) as a long polycistronic RNA which then undergoes posttranscriptional processing to produce the mature rRNAs. This occurs by the action of a set of endoribonucleases and exoribonucleases, some of which have been identified and characterized (Bollenbach et al., 2005, 2007). In some other cases, it is not entirely clear whether the defects in plastid rRNA processing observed in loss-of-function mutants are a primary or a secondary consequence of the gene knockout (e.g. Bellaoui et al., 2003).

All four rRNA species represent essential components of the plastid translational apparatus (Scharff and Bock, 2014). A peculiarity of the plastid 23S rRNA is its fragmentation into three distinct pieces in vivo, due to the so-called ‘hidden break’ processing (Nishimura et al., 2010). The efficiency of ‘hidden break’ processing is known to be dependent on environmental factors and developmental cues (Rosner et al., 1974), but the possible functional relevance of this fragmentation is unclear.

As in bacteria, the assembly of ribosomes in plastids depends on auxiliary proteins that aid rRNA folding and association with ribosomal proteins. Although a number of such ribosome biogenesis factors have recently been identified (e.g. Bellaoui et al., 2003; Bang et al., 2012; Lee et al., 2013; Fristedt et al., 2014), the precise molecular mode of action of most of them currently remains unknown.

Ribosomal Proteins

Identification of the complete set of plastid ribosomal proteins revealed that the protein composition of the plastid ribosome is very similar to that of the Escherichia coli ribosome (Tables 1 and 2). However, five additional so-called plastid-specific ribosomal proteins (PSRPs) were identified (Yamaguchi and Subramanian, 2000; Yamaguchi et al., 2000) that lack homologs in E. coli. The chloroplast 30S ribosomal subunit comprises a total of 24 proteins (Table 2) of which 21 are orthologs of E. coli 30S ribosomal proteins (S1-S21) and three are specific to chloroplast ribosomes (PSPR2, PSRP3, PSRP4; Yamaguchi et al., 2000). In the 50S subunit, 31 out of 33 ribosomal proteins have orthologs in E. coli (L1-L6, L9-L24, L27-L29, and L31-L36; Table 1) and two proteins are specific to chloroplasts (PSRP5 and PSRP6; Yamaguchi and Subramanian, 2000).

Table 1.

Ribosomal Proteins of the Large (50S) Subunit of the Chloroplast Ribosome and Their Genes in Bacteria (Escherichia coli) and Seed Plants (http://ppdb.tc.cornell.edu/dbsearch/subproteome.aspx).

Gene E. coli Reference A.t. gene Seed plants Reference
rpl1 Non-essential Baba et al., 2006 At3g63490 Essential Bryant et al., 2011; Romani et al., 2012
rpl2 Essential Shoji et al., 2011 Atcg01310
Atcg00830
NA
rpl3 Essential Shoji et al., 2011 At2g43030 Essential (NC) Unpublished
rpl4 Essential Shoji et al., 2011 At1g07320 Essential Romani et al., 2012
rpl5 Essential Shoji et al., 2011 At4g01310 NA
rpl6 Essential Shoji et al., 2011 At1g05190 Essential www.seedgenes.org
rpl9 Non-essential Baba et al., 2006 At3g44890 NA
rpl10 Essential Shoji et al., 2011 At5g13510 Essential (NC) Bryant et al., 2011
rpl11 Non-essential Baba et al., 2006 At1g32990 Non-essential Pesaresi et al., 2001
rpl12 Essential Shoji et al., 2011 At3g27830
At3g27850
NA
rpl13 Essential Shoji et al., 2011 At1g78630 Essential (NC) Bryant et al., 2011
rpl14 Essential Shoji et al., 2011 Atcg00780 NA
rpl15 Non-essential Shoji et al., 2011* At3g25920 NA
rpl16 Essential Shoji et al., 2011 Atcg00790 NA
rpl17 Essential Shoji et al., 2011 At3g54210 NA
rpl18 Essential Shoji et al., 2011 At1g48350 Essential (NC) Bryant et al., 2011
rpl19 Essential Shoji et al., 2011 At5g47190
At4g17560
NA
rpl20 Essential Shoji et al., 2011 Atcg00660 Essential Rogalski et al., 2008b
rpl21 Non-essential Shoji et al., 2011* At1g35680 Essential Yin et al., 2012
rpl22 Essential Shoji et al., 2011 Atcg00810 Essential Fleischmann et al., 2011
rpl23 Essential Shoji et al., 2011 Atcg01300
Atcg00840
Essential Fleischmann et al., 2011
rpl24 Non-essential Shoji et al., 2011* At5g54600 Non-essential Tiller et al., 2012
rpl25 Non-essential Baba et al., 2006
rpl27 Non-essential Shoji et al., 2011* At5g40950 Essential Romani et al., 2012
rpl28 Non-essential Dabbs, 1991* At2g33450 Essential Romani et al., 2012
rpl29 Non-essential Shoji et al., 2011* At5g65220 NA
rpl30 Non-essential Shoji et al., 2011*
rpl31 Non-essential Baba et al., 2006 At1g75350 Essential (NC) Bryant et al., 2011
rpl32 Non-essential Baba et al., 2006 Atcg01020 Essential Fleischmann et al., 2011
rpl33 Non-essential Baba et al., 2006 Atcg00640 Non-essential Rogalski et al., 2008b
rpl34 Non-essential Shoji et al., 2011* At1g29070 NA
rpl35 Essential Baba et al., 2006 At2g24090 Essential Romani et al., 2012
rpl36 Non-essential Baba et al., 2006 Atcg00760 Non-essential Fleischmann et al., 2011
psrp5 At3g56910 Knockdown viable Tiller et al., 2012
psrp6 At5g17870 Non-essential Tiller et al., 2012

Reverse genetic data for plants are from Arabidopsis (nucleus-encoded genes) and tobacco (plastid-encoded genes). Plastid genes are indicated in bold. All gene identifier codes are from Arabidopsis. For genes in the tobacco plastid genome, see GenBank accession Z00044.2. A.t., Arabidopsis thaliana; NA, not analyzed; NC, not yet confirmed by independent alleles or genetic complementation; *, contradictory result published earlier; unpublished, unpublished data from the authors’ laboratory; –, gene absent (from E. coli or plants).

Table 2.

Ribosomal Proteins of the Small (30S) Subunit of the Chloroplast Ribosome and Their Genes in Bacteria (Escherichia coli) and Seed Plants (http://ppdb.tc.cornell.edu/dbsearch/subproteome.aspx).

Gene E. coli Reference A.t. gene Seed plants Reference
rps1 Essential Shoji et al., 2011 At5g30510 Knockdown viable Romani et al., 2012
rps2 Essential Shoji et al., 2011 Atcg00160 Essential Rogalski et al., 2008b
rps3 Essential Shoji et al., 2011 Atcg00800 Essential Fleischmann et al., 2011
rps4 Essential Shoji et al., 2011 Atcg00380 Essential Rogalski et al., 2008b
rps5 Essential Shoji et al., 2011 At2g33800 Essential Bryant et al., 2011
rps6 Non-essential Bubunenko et al., 2007 At1g64510 NA
rps7 Essential Shoji et al., 2011 Atcg00900
Atcg01240
NA
rps8 Essential Shoji et al., 2011 Atcg00770 NA
rps9 Non-essential Shoji et al., 2011* At1g74970 Essential (NC) Bryant et al., 2011; Ma and Dooner, 2004
rps10 Essential Shoji et al., 2011 At3g13120 NA
rps11 Essential Shoji et al., 2011 Atcg00750 NA
rps12 Essential Shoji et al., 2011 Atcg00905
Atcg01230
Atcg00065
NA
rps13 Non-essential Bubunenko et al., 2007* At5g14320 Essential Bryant et al., 2011
rps14 Essential Shoji et al., 2011 Atcg00330 Essential Ahlert et al., 2003
rps15 Non-essential Bubunenko et al., 2007 Atcg01120 Non-essential Fleischmann et al., 2011
rps16 Essential Shoji et al., 2011 Atcg00050 Essential Fleischmann et al., 2011
rps17 Non-essential Shoji et al., 2011* At1g79850 Non-essential Schultes et al., 2000; Romani et al., 2012
rps18 Essential Shoji et al., 2011 Atcg00650 Essential Rogalski et al., 2006
rps19 Essential Shoji et al., 2011 Atcg00820 NA
rps20 Non-essential Bubunenko et al., 2007 At3g15190 Essential Romani et al., 2012; Gong et al., 2013
rps21 Non-essential Baba et al., 2006 At3g27160 Non-essential Morita-Yamamuro et al., 2004
psrp2 At3g52150 Knockdown viable Tiller et al., 2012
psrp3 At1g68590 Non-essential Tiller et al., 2012
psrp4 At2g38140 Knockdown viable Tiller et al., 2012

Reverse genetic data for plants are mainly from Arabidopsis (nucleus-encoded genes) and tobacco (plastid-encoded genes). Plastid genes are indicated in bold. All gene identifier codes are from Arabidopsis. For genes in the tobacco plastid genome, see GenBank accession Z00044.2. A.t., Arabidopsis thaliana; NA, not analyzed; NC, not yet confirmed by independent alleles or genetic complementation; *, contradictory result published earlier; –, gene absent (from E. coli or plants).

Genome sequencing revealed that the genes encoding plastid ribosomal proteins are distributed between the plastid and the nuclear genomes. Twelve proteins of the small subunit are encoded in the nuclear genome, whereas 12 are encoded in the plastid genome (Yamaguchi et al., 2000). By contrast, the majority of the genes encoding ribosomal proteins of the large subunit (24 genes) were transferred to the nucleus and only nine ribosomal protein genes remain in the plastid genome (Yamaguchi and Subramanian, 2000).

Some chloroplast ribosomal proteins are larger than their bacterial counterparts due to the presence of N-terminal and/or C-terminal extensions, the functional significance of which is not clear (Yamaguchi and Subramanian, 2000; Yamaguchi et al., 2000). These size expansions and the presence of the PSRPs result in a significant alteration of the protein: RNA ratio of the ribosome, which is 2:3 in the chloroplast, but 1:3 in E. coli.

The rRNAs are only stable if integrated into ribosomal subunits. Therefore, rRNA accumulation can serve as a proxy for chloroplast ribosome content. In addition, determining the abundance of individual rRNA species provides information about ribosomal subunit accumulation (Walter et al., 2010). For example, alterations in the ratio of 16S:23S rRNA are indicative of disturbances in ribosome assembly or stability that are specific to the 30S or 50S subunit. This indirect method of ribosome quantification has been employed in Arabidopsis and tobacco for the identification and characterization of mutants affected in chloroplast translation and ribosome biogenesis (Walter et al., 2010; Fleischmann et al., 2011; Tiller et al., 2012; Fristedt et al., 2014).

Comprehensive data concerning the essentiality of ribosomal proteins are available for E. coli (Dabbs, 1991; Baba et al., 2006; Shoji et al., 2011). Based on the strikingly similar overall composition of chloroplast and bacterial ribosomes, it seemed reasonable to assume that also the essentiality and non-essentiality of ribosomal proteins would be conserved between bacteria and plastids. However, recent studies indicate that this is not always the case (Tables 1 and 2).

The function of the nucleus-encoded plastid ribosomal proteins has been studied mainly in the model plant Arabidopsis by using publicly available T-DNA insertion lines (Pesaresi et al., 2001; Morita-Yamamuro et al., 2004; Bryant et al., 2011; Tiller et al., 2012; Romani et al., 2012), mutants generated by RNAi approaches (Tiller et al., 2012), or mutants identified in EMS mutagenesis screens (Yin et al., 2012). Due to the lack of a chloroplast transformation system for Arabidopsis, the essentiality of plastid-encoded ribosomal proteins has been exclusively investigated in tobacco (Nicotiana tabacum). These studies revealed that, similar to the results obtained in E. coli, the ribosomal protein genes rpl20, rpl22, rpl23, rps2, rps3, rps4, rps14, rps16, and rps18 are essential genes in tobacco plastids (Tables 1 and 2). Targeted inactivation of these genes was attempted by insertional mutagenesis taking advantage of the active homologous recombination system present in chloroplasts. Disruption of an essential ribosomal protein gene with a selectable marker gene for chloroplast transformation that confers resistance to the aminoglycoside antibiotic spectinomycin produces plastid transformants (transplastomic lines) that remain heteroplasmic (i.e. maintain a mix of transformed and wild-type copies of the plastid DNA). This is indicative of a balancing selection in which the wild-type genome copies provide the essential ribosomal protein and the transformed genome copies provide the enzyme detoxifying the selecting antibiotic. When such heteroplasmic plants are grown in the absence of the antibiotic, they show a characteristic leaf-loss phenotype (Figure 1). Due to random genome segregation into homoplasmy, cell lineages that entirely lack the essential plastid gene die. This leads to misshapen leaves that can lack large sectors of their lamina (Figure 1; Ahlert et al., 2003; Rogalski et al., 2006, 2008b; Fleischmann et al., 2011).

Figure 1.

Figure 1.

Typical Leaf Phenotypes of Transplastomic Plants Harboring a Knockout Allele for an Essential Component of the Translational Apparatus.

Leaves of transplastomic tobacco plants transformed with a knockout construct for the essential plastid gene rps18 (encoding plastid ribosomal protein S18; Rogalski et al., 2006; Table 2) are shown. The plastid transformants are heteroplasmic and, in the absence of antibiotic selection, the plants randomly segregate into homoplasmy for the wild-type plastid genome or homoplasmy for the transplastome. Homoplasmy for the rps18 knockout allele is lethal at the cellular level and results in loss of cell proliferation. Death of cell lineages during leaf development produces aberrantly shaped leaves that lack individual sectors or, in extreme cases, nearly the entire leaf blade (Ahlert et al., 2003). Scale bar = 2 cm.

Three plastid-encoded ribosomal protein genes turned out to be non-essential in reverse genetic studies: rpl33, rpl36, and rps15 (Rogalski et al., 2008b; Fleischmann et al., 2011). The homologous genes were also shown to be dispensable in E. coli (Tables 1 and 2). While Δrpl33 and Δrps15 transplastomic knockout mutants show wild-type-like growth under standard growth conditions, they are more sensitive to chilling stress than the wild-type, suggesting that both proteins are required for maintaining sufficiently high chloroplast translational capacity in the cold (Rogalski et al., 2008b; Fleischmann et al., 2011). By contrast, the loss of Rpl36 causes a severe mutant phenotype. Plants grow extremely slowly, display severe pigment deficiency, and show strongly altered leaf morphology (Figure 2; Fleischmann et al., 2011; see below).

Figure 2.

Figure 2.

Knockout of a Non-Essential Plastid Gene Involved in Translation.

The large ribosomal subunit protein L36 is not essential for plastid translation and, therefore, homoplasmic knockout mutants can be obtained (Fleischmann et al., 2011). Tobacco Δrpl36 plants grow autotrophically in soil, but suffer from severe photo-oxidative damage due to low levels of plastid protein biosynthesis which results in low photosynthetic activity. Moreover, the mutants display striking alterations in leaf morphology and plant architecture (Fleischmann et al., 2011).

(A) A young Δrpl36 plant two month after transfer from in vitro culture to the greenhouse. Note the much narrower leaf blade compared to the wild-type shown in (B). Scale bar = 2 cm.

(B) Leaf shape of a young wild-type plant at approximately the same developmental stage as the Δrpl36 plant shown in (A). Scale bar = 2 cm.

(C) A flowering wild-type plant (after 9 weeks of growth in the greenhouse). Scale bar = 10 cm.

(D) A Δrpl36 plant after 1.3 years of growth in the greenhouse. Note the narrow leaves, the extensive atypical shoot branching (indicative of reduced apical dominance), and the lack of floral induction. Scale bar = 10 cm.

Contradictory results were obtained for rpl32. While Rpl32-deficient E. coli cells were reported to be viable (Baba et al., 2006), attempts to knockout the rpl32 gene in tobacco strongly suggested its essentiality in plastids (Fleischmann et al., 2011). Re-analysis of the E. coli rpl32 knockout strain from the Keio collection revealed that it contained an intact rpl32 gene, raising the possibility that rpl32 is also essential in E. coli (Fleischmann et al., 2011).

Much less is known about the essentiality of the nucleus-encoded subunits of the chloroplast ribosome. Results from the proteins studied to date by reverse genetic approaches do not suggest a strict conservation of essentiality and non-essentiality between E. coli and plastids (Romani et al., 2012). However, data for some E. coli proteins are controversial (Tables 1 and 2) and the rpl32 case suggests that the results from high-throughput knockout projects may need to be treated with caution. rpl11, rpl24, rps17, and rps21 null mutants are viable in Arabidopsis, which is in line with L11, L24, S17, and S21 being non-essential for the assembly of functional ribosomes in E. coli (Tables 1 and 2). Knockout plants for these ribosomal protein genes display retarded growth, reduced activity of plastid protein biosynthesis, and impaired photosynthesis (Pesaresi et al., 2001; Morita-Yamamuro et al., 2004; Romani et al., 2012; Tiller et al., 2012).

In agreement with data obtained for E. coli, the nucleus-encoded L4, L6, L35, and S5 proteins are essential in Arabidopsis (www.seedgenes.org; Tables 1 and 2). The genes for L3, L10, L13, and L18 are also likely to be essential in both E. coli and Arabidopsis (Table 1), but more rigorous experimental confirmation is still needed. A third category of nucleus-encoded plastid ribosomal proteins comprises L1 and S20, which were shown to be essential in Arabidopsis but are non-essential in E. coli (Tables 1 and 2). It is currently not clear whether L21, L27, L28, and S13 also fall in this category, because conflicting data have been reported for E. coli (Tables 1 and 2).

The initially six plastid-specific ribosomal proteins (PSRP1-6) were suggested to play structural roles and regulate chloroplast translation in response to light (Yamaguchi and Subramanian, 2003; Manuell et al., 2007). The occurrence of PSRP2, PSRP3, PSRP4, PSRP5, and PSRP6 in stoichiometric amounts with the classical ribosomal proteins indicated that they could represent bona fide ribosomal proteins. Recently, PSRP1 has been proven to be neither a ribosomal protein nor plastid-specific. Instead, it appears to be a functional homolog of the E. coli cold-shock protein pY that acts as a ribosome-binding translation factor (Sharma et al., 2007, 2010). Cryo-electron microscopic studies localized PSRP2, PSRP3, PSRP4, and PSRP5 in the three-dimensional structure of the chloroplast ribosome (Manuell et al., 2007; Sharma et al., 2007) and suggest that these PSRPs might play structural roles, perhaps by compensating for the loss of specific rRNA sequences and/or secondary structural elements. PSRP6 might be only loosely associated with the plastid ribosome (Sharma et al., 2007).

In a recent study, a reverse genetic approach was taken in Arabidopsis to elucidate whether the remaining five PSRPs (PSRP2-6) are required for chloroplast translation or ribosome assembly (Tiller et al., 2012). Mutants were analyzed by describing their phenotypes, measuring their photosynthetic capacity, and analyzing their chloroplast translation efficiency. To determine the impact of reduced or abolished PSRP expression on ribosome assembly, ribosomal RNAs were precisely quantified. The data obtained suggest that PSRP3 and PSRP4 in the 30S subunit as well as PSRP5 in the 50S subunit play structural roles in the ribosome and thus qualify as genuine ribosomal proteins. Their down-regulation leads to decreased accumulation of the 30S or 50S subunits of the plastid ribosome. Although assigned as a protein of the small ribosomal subunit, PSRP3 also affects the accumulation of the 50S ribosomal subunit, possibly suggesting a dual role for PSRP3 in ribosome biogenesis and/or stability. By contrast, plants with reduced amounts of PSRP2 (in the 30S subunit) and plants without PSRP6 (in the 50S subunit) display no visible or measurable phenotypes, suggesting that they are non-essential proteins and dispensable for ribosome biogenesis and translation at least under standard conditions. PSRP2 contains two RNA-binding domains and recently was proposed to possess RNA chaperone activity and function as negative regulator in seed germination and abiotic stress responses (Yamaguchi and Subramanian, 2003; Xu et al., 2013).

The set of ribosomal protein genes that is retained in the chloroplast genome is highly conserved across seed plants. Known exceptions include rpl32 which was transferred to the nuclear genome in poplar (Ueda et al., 2007), rpl22 which was transferred to the nucleus in legumes (Gantt et al., 1991), and rpl23 which is a pseudogene in the spinach plastid genome and whose gene product was replaced by a eukaryotic L23 protein that is encoded in the nucleus and imported into the chloroplast (Bubunenko et al., 1994).

Transfer RNAs

The genetic code comprises 64 triplets, 61 of which specify the 20 different proteinogenic amino acids. Due to the relaxed base pairing between the third codon position in the mRNA and the first position of the anticodon in the tRNA, fewer than 61 tRNA species are needed to read all 61 triplets. This was early recognized by Francis Crick and has become known as the wobble hypothesis (Crick, 1966). According to Crick’s wobble rules, 32 tRNA species should constitute the minimum set required for translation. The plastid genome of most seed plants encodes only 30 tRNA genes (Sugiura and Wakasugi, 1989; Alkatib et al., 2012b). In the absence of experimental evidence for the import of nucleus-encoded tRNAs into plastids, this raised the question how chloroplasts synthesize their proteins with fewer than 32 tRNAs.

A number of recent studies addressed this issue by employing reverse genetics in the model system tobacco. Systematic knockout analysis of individual tRNA genes revealed that a process called superwobbling facilitates translation with reduced tRNA sets (Rogalski et al., 2008a; Alkatib et al., 2012b). Superwobbling refers to the capability of a single tRNA species to read an entire codon box. (A codon box is defined as set of four codons sharing the first two nucleotides.) According to the classical wobble rules, at least two tRNA isoacceptors are needed to read all four triplets of a codon box. For example, the four glycine codons GGC, GGU, GGA, and GGG are decoded by two tRNAs: tRNA-Gly(GCC) and tRNA-Gly(UCC). While tRNA-Gly(GCC) reads GGC (by Watson–Crick base pairing) and GGU triplets (by wobbling), tRNA-Gly(UCC) recognizes GGA (by Watson–Crick base pairing) and GGG codons (by wobbling). However, knockout of the trnG-GCC gene in the tobacco plastid genome demonstrated that translation is sustained in the absence of tRNA-Gly(GCC). This suggests that a single glycine tRNA, tRNA-Gly(UCC), is sufficient to read all four glycine triplets. Apparently, the (unmodified) uridine in the first position of the anticodon of tRNA-Gly(UCC) can superwobble by engaging in base-pairing interactions with all four nucleotides (Rogalski et al., 2008a). Interestingly, the trnG-UCC gene encoding tRNA-Gly(UCC) is essential and tRNA-Gly(GCC) cannot read all four glycine codons. This is in agreement with theoretical considerations (Crick, 1966) that, for steric reasons, the small pyrimidine base U can base pair weakly with U and C, but the bulky purine base G cannot base pair with A and G. Superwobbling provides a straightforward explanation for the ‘missing tRNAs’ in chloroplast genomes (Alkatib et al., 2012b). It, however, cannot explain all cases of missing tRNA genes in the highly reduced plastid genomes of non-photosynthetic holoparasitic plants (Morden et al., 1991; Delannoy et al., 2011; Krause, 2011) and how these parasites perform protein biosynthesis in their plastid compartment currently remains a mystery.

Systematic reverse genetic analysis of tRNA genes in the tobacco plastome demonstrated that superwobbling occurs in all codon boxes where it is theoretically possible (Rogalski et al., 2008a; Alkatib et al., 2012a,b). These studies also defined the minimum set of tRNA species that is required to sustain protein biosynthesis. Upon maximum use of wobbling and superwobbling, the minimum tRNA set comprises 25 tRNA species (Alkatib et al., 2012b). However, this minimum set is only rarely found in nature because, in most cases, superwobbling appears to be considerably less efficient than conventional base pairing and wobbling. Consequently, most knockouts of non-essential plastid tRNA genes displayed mutant phenotypes characterized by reduced photosynthetic activity and retarded growth, as caused by their lowered levels of plastid translation (Rogalski et al., 2008a; Alkatib et al., 2012b). The phenotypes of mutants for essential plastid tRNA genes are very similar to those described above for essential plastid genes encoding ribosomal proteins (Alkatib et al., 2012a,b).

As in all other biological systems, plastid tRNAs are transcribed as precursors, undergo posttranscriptional processing (Vogel et al., 1997; Gutmann et al., 2012), and are subject to extensive nucleotide modifications, which are thought to improve decoding accuracy or influence tRNA folding and/or stability (Delannoy et al., 2009; Karcher and Bock, 2009).

PLASTID TRANSLATION AND PLANT ANATOMY

Characterizations of plants with defects in chloroplast gene expression revealed the importance of plastids for leaf architecture (Pyke et al., 2000; Wang et al., 2000; Wycliffe et al., 2005). The leaf architecture is defined by mesophyll and bundle sheath cells enclosed by a layer of epidermal cells on both sides of the leaf (i.e. the adaxial side and the abaxial side). The mesophyll is composed of palisade cells and more loosely packed spongy mesophyll cells. The cylindrically shaped palisade cells reside below the adaxial epidermis and are major contributors to the plant’s photosynthetic performance. The bundle sheath cells are embedded in the mesophyll and transport water, minerals, sugars, and other organic compounds. Analysis of mutants in plastid gene expression, including translation mutants, suggested a tight connection between chloroplast function and palisade cell development. Many mutants in plastid gene expression show not only defects in chloroplast biogenesis, but also disturbed palisade parenchyma development (Chatterjee et al., 1996; Wang et al., 2000; Wycliffe et al., 2005; Tiller et al., 2012).

Almost all levels of plastid gene expression have been implicated in the differentiation of palisade cells, including transcription (Hricová et al., 2006), RNA maturation (Reiter et al., 1994; Meurer et al., 1998a), and translation (Tiller et al., 2012). For example, defective palisade cell development has been reported in a wide spectrum of translation mutants, including mutants affected in components of the ribosome (Tiller et al., 2012), factors involved in ribosome biogenesis (Bang et al., 2012), and translation factors (e.g. elongation factor G and RRF; Ruppel and Hangarter, 2007; Wang et al., 2010). This suggests that sufficient translational capacity in the plastid (rather than specific components of the translational apparatus) is required for proper formation of the palisade parenchyma. This interpretation has gained further support from studies using specific inhibitors of chloroplast protein biosynthesis (Pyke et al., 2000).

It was suggested that the developmental and/or metabolic status of the plastids regulates cell differentiation and leaf morphogenesis by plastid-to-nucleus communication (Streatfield et al., 1999; Rodermel, 2001; Naested et al., 2004; Tan et al., 2008). Plastid retrograde signaling enables the chloroplast to communicate its functional and metabolic status to the host cell and to adjust nuclear gene expression according to the needs of the organelles (for recent reviews, see, e.g. Lepistö et al., 2012; Barajas-López et al., 2013; Estavillo et al., 2013; Jarvis and López-Juéz, 2013). With the currently available data, it is difficult to judge whether or not the defects in palisade cell development caused by impaired plastid translation represent a secondary consequence of disturbed chloroplast biogenesis. Preliminary evidence for the possibility to uncouple defects in plastid translation from impaired chloroplast biogenesis has recently come from the analysis of a set of plastid ribosomal protein mutants. When grown under comparable conditions, different mutants showed varying defects in palisade cell differentiation which, however, did not correlate well with the severity of the mutant phenotype (Tiller et al., 2012).

How does the protein biosynthesis capacity of the plastid affect mesophyll development? It seems conceivable that the deficiency in a specific plastid genome-encoded protein is responsible for the effect on palisade parenchyma development upon impaired plastid translation. Alternatively, the reduced expression capacity of the chloroplast could be sensed and transmitted by retrograde signaling to the nucleus, where it alters the gene expression program required for palisade cell differentiation. Whether the proposed plastid retrograde signal controlling mesophyll development is distinct from the classical (GUN-type) retrograde signaling remains to be clarified. Expression of LHCB, a favored marker gene of retrograde signaling (Susek et al., 1993; Sullivan and Gray, 1999), was investigated in some of the plastid gene expression mutants showing altered mesophyll organization. While LHCB expression appears to be unaffected in some mutants (e.g. in the pac1 mutant which is defective in plastid mRNA maturation; Reiter et al., 1994; Meurer et al., 1998a), it was found to be reduced in others (Hricová et al., 2006; reviewed, e.g., in Rodermel, 2001). This may provide preliminary evidence for the existence of a GUN-independent communication pathway influencing leaf anatomy, which vice versa is also supported by deregulated LHCB expression in the absence of disturbed palisade parenchyma development in retrograde signaling mutants, such as gun1 (Rodermel, 2001).

PLASTID TRANSLATION AND PLANT MORPHOLOGY

Plastid Influence on Leaf Development in Evening Primroses

Early genetic studies in evening primroses (genus Oenothera; Onagraceae) provided evidence for the genotype of the plastid controlling leaf shape (Schwemmle, 1938, 1941, 1943). This discovery was facilitated by three salient features of the evening primrose system: (1) biparental chloroplast inheritance, (2) crossability among species and formation of fertile interspecific hybrids, and (3) a genetic phenomenon referred to as permanent translocation heterozygosity (Rauwolf et al., 2008; Greiner and Bock, 2013). These features make it relatively straightforward to produce interspecific hybrids with identical nuclear genomes but different plastid genomes. Performing interspecific crosses between Oenothera odorata and Oenothera berteriana, German botanist Julius Schwemmle was the first to realize that the leaf shape in evening primroses co-segregates with the genotype of the plastid. Leaves of Oe. berteriana are wider and more serrated than Oe. odorata leaves. Hybrids of Oe. odorata and Oe. berteriana with berteriana plastids have berteriana-like (wider and more strongly toothed) leaves, hybrid plants with odorata plastids have odorata-like (narrower and less toothed) leaves. This work clearly established that the genotype of the plastid determines the shape of evening primrose leaves (Schwemmle, 1938, 1941, 1943; Hagemann, 1964), which at that time was referred to as an ‘extraplastidic trait’, because it was unrelated to leaf pigmentation and photosynthesis.

Plastid Translation and Leaf Development in Tobacco

Investigation of chloroplast gene expression by transplastomic approaches in tobacco revealed that plastid protein synthesis is absolutely required for plant development (Ahlert et al., 2003; Figure 2). Leaf and flower development are dependent on the expression of plastid genes in that loss of translation results in arrested cell division. This leads to a characteristic phenotype in which the death of entire cell lineages causes severely misshapen organs (see above; Figure 1). This phenotype was observed not only with knockout alleles for essential components of the plastid translational apparatus (Tables 1 and 2; Rogalski et al., 2008a; Alkatib et al., 2012b), but was also seen in heteroplasmic knockout mutants for other essential chloroplast genes, such as accD (encoding the D subunit of acetyl-CoA carboxylase, an enzyme required for fatty acid biosynthesis; Kode et al., 2005), clpP (encoding an essential subunit of the Clp protease; Shikanai et al., 2001), and the two large open reading frames ycf1 and ycf2 (Drescher et al., 2000). This raises the possibility that plastid translation is only essential because it is required to express the few essential plastid genes that function outside of plastid gene expression.

In addition to the essentiality of plastid translation for leaf formation, the translational activity in plastids also affects leaf shape. For example, loss of the (non-essential) gene for tRNA-Gly(GCC) strongly reduces plastid translational activity and, interestingly, also causes an altered shape of the cotyledons (which are slender in the mutant compared to round-shaped cotyledons in the wild-type; Rogalski et al., 2008a). A similar narrow leaf phenotype was observed in true leaves of tobacco plants carrying a knockout allele for the (non-essential) plastid ribosomal protein gene rpl36 (Fleischmann et al., 2011; Figure 2). Although both the trnG(GCC) and the rpl36 mutant also have a photosynthetic phenotype, reduced photosynthetic activity cannot cause the alterations in leaf shape, because none of the many photosynthesis mutants (including many mutants that entirely lack photosynthesis) has been reported to display leaf shape phenotypes (Hager et al., 1999, 2002; Krech et al., 2012).

The altered leaf shape seen in tobacco mutants with reduced plastid translational activity is reminiscent of the control of leaf shape by the plastid genotype in evening primroses (Schwemmle, 1938, 1941, 1943). However, whether or not there is a direct mechanistic link between these two phenomena (in that altered translation of a plastid gene causes the phenotypic difference between Oe. odorata and Oe. berteriana leaves; see above) remains to be established.

The size and shape of leaves are pivotal factors determining plant performance and photosynthetic activity under ever-changing environmental conditions. Leaf morphology influences key photosynthetic parameters, such as the efficiency of light absorption and gas exchange. Some aspects of leaf morphology (e.g. leaf area, leaf thickness, and leaf index, i.e. the ratio of leaf length to leaf width) respond to environmental cues, including light intensity, growth temperature, and water availability. It seems tempting to speculate that the altered leaf morphology upon impaired plastid translation is part of such an adaption mechanism. In this scenario, changes in the environmental conditions induce changes in the translational activity of the chloroplast which in turn generate a signal to modify leaf morphology. It is well established that the translational apparatus of plastids responds sensitively to abiotic stress, especially temperature stress (Hanson and Bogorad, 1978; Bünger and Feierabend, 1980; Hopkins and Elfman, 1984; Xu et al., 2013). Since reduced chloroplast translation has a direct impact on photosynthetic activity (Fleischmann et al., 2011; Tiller et al., 2012), it seems conceivable that low chloroplast translational activity generates a plastid signal that causes the development of more narrow leaves that absorb less light energy, thereby avoiding photo-oxidative damage.

Altered Leaf Development in Arabidopsis Mutants with Defects in Plastid Translation

The impact of plastid protein biosynthesis on leaf development was also confirmed by the identification and characterization of plastid gene expression mutants in the model plant Arabidopsis thaliana. As with the effects of plastid gene expression on leaf anatomy, mutations in any gene expression processes upstream of plastid translation are likely to also affect translation. For example, general defects in transcription or RNA processing are likely to have an impact on rRNAs, tRNAs, and/or mRNAs for ribosomal proteins and, therefore, unavoidably will result in lower levels of plastid translation. Consequently, mutants in plastid RNA metabolism and plastid translation mutants can have very similar phenotypes. For example, the Arabidopsis mutants scabra3 (defective in the gene for the nuclear-encoded plastid RNA polymerase RpoTp; Hricová et al., 2006) and hfp108–1 (defective in the chloroplast RRF; Wang et al., 2010) both have aberrantly shaped, strongly serrated leaves. A similar phenotype was also observed in a mutant defective in chloroplast ribosome biogenesis and, interestingly, correlated with reduced amounts of the phytohormone abscisic acid (Lee et al., 2013). In view of the crucial role of abscisic acid in leaf development in other systems (Chen et al., 2011; Wanke, 2011), this observation certainly warrants further investigation.

Since the functions of nearly all plastid genome-encoded gene products have been elucidated (Scharff and Bock, 2014) and none of them appears to be directly involved in leaf development, it seems reasonable to assume that the effect of plastid translation on leaf development is the result of retrograde signaling. This, however, remains to be firmly established. Also, the signaling molecule(s) and nuclear target genes of such a retrograde communication pathway will need to be identified. The development of the leaf blade is dependent on polar cell proliferation and polar cell expansion (reviewed, e.g., in Tsukaya, 1995; Sylvester et al., 1996; Byrne, 2012; Gonzalez et al., 2012). Leaf growth by cell proliferation is influenced by a number of genes in Arabidopsis, including the ROTUNDIFOLIA and ANGUSTIFOLIA genes, among others. While ANGUSTIFOLIA gene products are required for leaf expansion in the lateral direction, ROTUNDIFOLIA proteins mediate expansion in the longitudinal direction (Kim and Cho, 2006; Kawade et al., 2010). Whether or not these two groups of genes represent targets of signaling from plastids with impaired translational activity is currently not known. A recent study has provided evidence for plastid retrograde signaling influencing the spatial expression patterns of key genes involved early in expansion of the leaf lamina (Tameshige et al., 2013). In leaf primordia, the expression of the gene FILAMENTOUS FLOWER (FIL) and the activity of a microRNA (miR165/166) are specific to the abaxial side. Upon inhibition of plastid translation, the spatial restriction of FIL expression and miR165/166 activity is retarded and the leaf lamina becomes narrow (Tameshige et al., 2013).

Interestingly, genetic screens for mutants with alterations in leaf shape have revealed not only genes involved in plastid translation (Moschopoulos et al., 2012), but also several genes for ribosomal proteins of the cytosolic (80S) ribosome (Pinon et al., 2008; Yao et al., 2008). Why mutations in some cytosolic ribosomal protein genes cause leaf phenotypes while mutations in many others do not is currently unknown. Ribosome heterogeneity in the cytosol or formation of aberrant ribosomes has been proposed as a possible explanation, but more work is needed to resolve this conundrum (Byrne, 2012; Horiguchi et al., 2012). Nonetheless, the striking mutant phenotypes caused by defects in both the cytosolic and the plastid protein biosynthesis machineries suggest an intimate relationship between translation in the cytosol and translation in the plastid compartment in controlling leaf shape. To elucidate the underlying molecular mechanisms and to determine the physiological significance of this regulation (and, in particular, its possible role in optimizing photosynthesis at the level of organ size and organ shape) represent major challenges for the future.

The Impact of Plastid Translation on Flower Development and Other Morphological Traits

Transplastomic experiments have established that plastid protein biosynthesis is also required for flower development (Ahlert et al., 2003). Consistently with this finding and similarly to the observed effects of the plastid genotype on leaf shape, early genetic evidence from interspecific crosses in evening primroses also suggested an influence of the plastid genome on flower morphology. Reciprocal crosses between Oe. odorata and Oe. berteriana revealed that the length of the hypanthium is under genetic control of the plastid (Schwemmle, 1938, 1941, 1943). Unfortunately, this potentially very interesting finding was never followed up on and its interpretation remained controversial (Rhoades, 1955; Hagemann, 1964). The same holds true for the reported quantitative effect of the plastid genotype on stem growth (Hagemann, 1964). Interestingly, a potentially related observation was made in transplastomic knockout mutants of the gene for ribosomal protein L36 (Fleischmann et al., 2011). These mutants display strongly reduced apical dominance leading to excessive stem branching, thus giving the plants a bushy appearance (Figure 2C and 2D). As apical dominance is mainly determined by concentration gradients of the phytohormone auxin, this finding could suggest that impaired plastid translation influences auxin biosynthesis and/or transport. This could occur either by a plastid genome-encoded gene product influencing the activity or turnover of one of the (nucleus-encoded) enzymes of auxin biosynthesis or, alternatively, via a retrograde signaling pathway controlling the expression level of nuclear genes involved in auxin synthesis, transport, or degradation.

In summary, although it has long been known that the fidelity of plastid gene expression influences gene expression in the nucleus (Börner et al., 1986; Hess et al., 1994) and much has been learnt about the underlying signaling pathways (Susek et al., 1993; Koussevitzky et al., 2007), only recently phenotypic traits have been firmly associated with the fidelity of plastid gene expression. It appears likely that the translational activity of the plastid can generate a retrograde signal that influences specific aspects of plant anatomy and morphology, but how this signal fits into the increasingly complex landscape of plastid-to-nucleus communication (Leister, 2012; Barajas-López et al., 2013) remains to be elucidated. Future work should be directed towards (1) the discovery of components of the signaling pathway that act downstream of plastid translation and (2) the identification of nuclear target genes that produce the ultimate morphological output of this pathway.

FUNDING

Work on plastid translation in the authors’ laboratory is supported by the Max Planck Society and by a grant from the Deutsche Forschungsgemeinschaft to R.B. (FOR 804; BO 1482/15–2).

ACKNOWLEDGMENTS

We thank Drs Tobias Fleischmann, Lars Scharff, and Marcelo Rogalski for providing photographs of mutant plants, and Dr. Rodrigo Caroca (all MPI-MP) for sharing unpublished data on ribosomal protein mutants. No conflict of interest declared.

REFERENCES

  1. Ahlert D., Ruf S., Bock R. (2003). Plastid protein synthesis is required for plant development in tobacco. Proc. Natl Acad. Sci. U S A. 100, 15730–15735 [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Albrecht V., Ingenfeld A., Apel K. (2006). Characterization of the snowy cotyledon 1 mutant of Arabidopsis thaliana: the impact of chloroplast elongation factor G on chloroplast development and plant vitality. Plant Mol. Biol. 60, 507–518 [DOI] [PubMed] [Google Scholar]
  3. Alkatib S., Fleischmann T.T., Scharff L.B., Bock R. (2012a). Evolutionary constraints on the plastid tRNA set decoding methionine and isoleucine. Nucleic Acids Res. 40, 6713–6724 [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Alkatib S., Scharff L.B., Rogalski M., Fleischmann T.T., Matthes A., Seeger S., Schöttler M.A., Ruf S., Bock R. (2012b). The contributions of wobbling and superwobbling to the reading of the genetic code. PLoS Genet. 8, e1003076. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Allison L.A., Maliga P. (1995). Light-responsive and transcription-enhancing elements regulate the plastid psbD core promoter. EMBO J. 14, 3721–3730 [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Allison L.A., Simon L.D., Maliga P. (1996). Deletion of rpoB reveals a second distinct transcription system in plastids of higher plants. EMBO J. 15, 2802–2809 [PMC free article] [PubMed] [Google Scholar]
  7. Baba T., Ara T., Hasegawa M., Takai Y., Okumura Y., Baba M., Datsenko K.A., Tomita M., Wanner B.L., Mori H. (2006). Construction of Escherichia coli K-12 in-frame, single-gene knockout mutants: the Keio collection. Mol. Syst. Biol. 2, 2006.0008, 21 February, 10.1038/msb4100050 [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Bang W.Y., Chen J., Jeong I.S., Kim S.W., Kim C.W., Jung H.S., Lee K.H., Kweon H.-S., Yoko I., Shiina T., et al. (2012). Functional characterization of ObgC in ribosome biogenesis during chloroplast development. Plant J. 71, 122–134 [DOI] [PubMed] [Google Scholar]
  9. Barajas-López J.d.D., Blanco N.E., Strand A. (2013). Plastid-to-nucleus communication, signals controlling the running of the plant cell. Biochim. Biophys. Acta. 1833, 425–437 [DOI] [PubMed] [Google Scholar]
  10. Barkan A. (1988). Proteins encoded by a complex chloroplast transcription unit are each translated from both monocistronic and polycistronic mRNAs. EMBO J. 7, 2637–2644 [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Barkan A. (1998). Approaches to investigating nuclear genes that function in chloroplast biogenesis in land plants. Methods Enzymol. 297, 38–57 [Google Scholar]
  12. Barkan A. (2011). Expression of plastid genes: organelle-specific elaborations on a prokaryotic scaffold. Plant Physiol. 155, 1520–1532 [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Beligni M.V., Yamaguchi K., Mayfield S.P. (2004a). Chloroplast elongation factor Ts pro-protein is an evolutionarily conserved fusion with the S1 domain-containing plastid-specific ribosomal protein-7. Plant Cell. 16, 3367–3369 [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Beligni M.V., Yamaguchi K., Mayfield S.P. (2004b). The translational apparatus of Chlamydomonas reinhardtii chloroplast. Photosynth. Res. 82, 315–325 [DOI] [PubMed] [Google Scholar]
  15. Bellaoui M., Keddie J.S., Gruissem W. (2003). DCL is a plant-specific protein required for plastid ribosomal RNA processing and embryo development. Plant Mol. Biol. 53, 531–543 [DOI] [PubMed] [Google Scholar]
  16. Bock R. (2000). Sense from nonsense: how the genetic information of chloroplasts is altered by RNA editing. Biochimie. 82, 549–557 [DOI] [PubMed] [Google Scholar]
  17. Bock R. (2007). Structure, function, and inheritance of plastid genomes. Top. Curr. Genet. 19, 29–63 [Google Scholar]
  18. Bock R., Timmis J.N. (2008). Reconstructing evolution: gene transfer from plastids to the nucleus. BioEssays. 30, 556–566 [DOI] [PubMed] [Google Scholar]
  19. Boeck R., Kolakofsky D. (1994). Positions +5 and +6 can be major determinants of the efficiency of non-AUG initiation codons for protein synthesis. EMBO J. 13, 3608–3617 [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Bollenbach T.J., Lange H., Gutierrez R., Erhardt M., Stern D.B., Gagliardi D. (2005). RNR1, a 3’-5’ exoribonuclease belonging to the RNR superfamily, catalyzes 3’ maturation of chloroplast ribosomal RNAs in Arabidopsis thaliana . Nucleic Acids Res. 33, 2751–2763 [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Bollenbach T.J., Schuster G., Portnoy V., Stern D.B. (2007). Processing, degradation, and polyadenylation of chloroplast transcripts. Top. Curr. Genet. 19, 175–211 [Google Scholar]
  22. Börner T., Mendel R.R., Schiemann J. (1986). Nitrate reductase is not accumulated in chloroplast-ribosome-deficient mutants of higher plants. Planta. 169, 202–207 [DOI] [PubMed] [Google Scholar]
  23. Bryant N., Lloyd J., Sweeney C., Myouga F., Meinke D. (2011). Identification of nuclear genes encoding chloroplast-localized proteins required for embryo development in Arabidopsis . Plant Physiol. 155, 1678–1689 [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Bubunenko M., Baker T., Court D.L. (2007). Essentiality of ribosomal and transcription antitermination proteins analyzed by systematic gene replacement in Escherichia coli . J. Bacteriol. 189, 2844–2853 [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Bubunenko M.G., Schmidt J., Subramanian A.R. (1994). Protein substitution in chloroplast ribosome evolution: a eukaryotic cytosolic protein has replaced its organelle homologue (L23) in spinach. J. Mol. Biol. 240, 28–41 [DOI] [PubMed] [Google Scholar]
  26. Bünger W., Feierabend J. (1980). Capacity for RNA synthesis in 70S ribosome-deficient plastids of heat-bleached rye leaves. Planta. 149, 163–169 [DOI] [PubMed] [Google Scholar]
  27. Byrne M.E. (2012). Making leaves. Curr. Opin. Plant Biol. 15, 24–30 [DOI] [PubMed] [Google Scholar]
  28. Chatterjee M., Sparvoli S., Edmunds C., Garosi P., Findlay K., Martin C. (1996). DAG, a gene required for chloroplast differentiation and palisade development in Antirrhinum majus . EMBO J. 15, 4194–4207 [PMC free article] [PubMed] [Google Scholar]
  29. Chen H.-C., Hwang S.-G., Chen S.-M., Shii C.-T., Cheng W.-H. (2011). ABA-mediated heterophylly is regulated by differential expression of 9-cis-epoxycarotenoid dioxygenase 3 in lilies. Plant Cell Physiol. 52, 1806–1821 [DOI] [PubMed] [Google Scholar]
  30. Choquet Y., Stern D.B., Wostrikoff K., Kuras R., Girard-Bascou J., Wollman F.-A. (1998). Translation of cytochrome f is autoregulated through the 5’ untranslated region of petA mRNA in Chlamydomonas reinhardtii . Proc. Natl Acad. Sci. U S A. 95, 4380–4385 [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Crick F.H.C. (1966). Codon-anticodon pairing: the wobble hypothesis. J. Mol. Biol. 19, 548–555 [DOI] [PubMed] [Google Scholar]
  32. Dabbs E.R. (1991). Mutants lacking individual ribosomal proteins as a tool to investigate ribosomal properties. Biochimie. 73, 639–645 [DOI] [PubMed] [Google Scholar]
  33. Delannoy E., Fujii S., des Francs-Small C.C., Brundrett M., Small I. (2011). Rampant gene loss in the underground orchid Rhizanthella gardneri highlights evolutionary constraints on plastid genomes. Mol. Biol. Evol. 28, 2077–2086 [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Delannoy E., Le Ret M., Faivre-Nitschke E., Estavillo G.M., Bergdoll M., Taylor N.L., Pogson B.J., Small I., Imbault P., Gualberto J.M. (2009). Arabidopsis tRNA adenosine deaminase arginine edits the wobble nucleotide of chloroplast tRNAArg(ACG) and is essential for efficient chloroplast translation. Plant Cell. 21, 2058–2071 [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Drechsel O., Bock R. (2010). Selection of Shine–Dalgarno sequences in plastids. Nucleic Acids Res. 39, 1427–1438 [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Drescher A., Ruf S., Calsa T., Jr., Carrer H., Bock R. (2000). The two largest chloroplast genome-encoded open reading frames of higher plants are essential genes. Plant J. 22, 97–104 [DOI] [PubMed] [Google Scholar]
  37. Eberhard S., Drapier D., Wollman F.-A. (2002). Searching limiting steps in the expression of chloroplast-encoded proteins: relations between gene copy number, transcription, transcript abundance and translation rate in the chloroplast of Chlamydomonas reinhardtii . Plant J. 31, 149–160 [DOI] [PubMed] [Google Scholar]
  38. Edwards K., Bedbrook J., Dyer T., Kössel H. (1981). 4.5S rRNA from Zea mays chloroplasts shows structural homology with the 3’ end of prokaryotic 23S rRNA. Biochem. Int. 2, 533–538 [Google Scholar]
  39. Estavillo G.M., Chan K.X., Phua S.Y., Pogson B.J. (2013). Reconsidering the nature and mode of action of metabolite retrograde signals from the chloroplast. Front. Plant Sci. 3, 300. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Fleischmann T.T., Scharff L.B., Alkatib S., Hasdorf S., Schöttler M.A., Bock R. (2011). Nonessential plastid-encoded ribosomal proteins in tobacco: a developmental role for plastid translation and implications for reductive genome evolution. Plant Cell. 23, 3137–3155 [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Fristedt R., Scharff L.B., Clarke C.A., Wang Q., Lin C., Merchant S.S., Bock R. (2014). RBF1, a plant homolog of the bacterial ribosome-binding factor RbfA, acts in processing of the chloroplast 16S ribosomal RNA. Plant Physiol. 164, 201–215 [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Gantt J.S., Baldauf S.L., Calie P.J., Weeden N.F., Palmer J.D. (1991). Transfer of rpl22 to the nucleus greatly preceded its loss from the chloroplast and involved the gain of an intron. EMBO J. 10, 3073–3078 [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Gong X., Jiang Q., Xu J., Zhang J., Teng S., Lin D., Dong Y. (2013). Disruption of the rice plastid ribosomal protein S20 leads to chloroplast developmental defects and seedling lethality. G3. 3, 1769–1777 [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Gonzalez N., Vanhaeren H., Inzé D. (2012). Leaf size control: complex coordination of cell division and expansion. Trends Plant Sci. 17, 332–340 [DOI] [PubMed] [Google Scholar]
  45. Gray M.W. (1993). Origin and evolution of organelle genomes. Curr. Opin. Genet. Dev. 3, 884–890 [DOI] [PubMed] [Google Scholar]
  46. Greiner S., Bock R. (2013). Tuning a ménage à trois: co-evolution and co-adaptation of nuclear and organellar genomes in plants. BioEssays. 35, 354–365 [DOI] [PubMed] [Google Scholar]
  47. Gutmann B., Gobert A., Giegé P. (2012). PRORP proteins support RNase P activity in both organelles and the nucleus in Arabidopsis . Genes Dev. 26, 1022–1027 [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Hagemann R. (1964). Plasmatische Vererbung: VEB G (Jena: Fischer Verlag; ). [Google Scholar]
  49. Hager M., Biehler K., Illerhaus J., Ruf S., Bock R. (1999). Targeted inactivation of the smallest plastid genome-encoded open reading frame reveals a novel and essential subunit of the cytochrome b6f complex. EMBO J. 18, 5834–5842 [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Hager M., Hermann M., Biehler K., Krieger-Liszkay A., Bock R. (2002). Lack of the small plastid-encoded PsbJ polypeptide results in a defective water-splitting apparatus of photosystem II, reduced photosystem I levels, and hypersensitivity to light. J. Biol. Chem. 277, 14031–14039 [DOI] [PubMed] [Google Scholar]
  51. Hanson M.R., Bogorad L. (1978). The ery-M2 group of Chlamydomonas reinhardii: cold-sensitive, erythromycin-resistant mutants deficient in chloroplast ribosomes. J. Gen. Microbiol. 105, 253–262 [Google Scholar]
  52. Hedtke B., Legen J., Weihe A., Herrmann R.G., Börner T. (2002). Six active phage-type RNA polymerase genes in Nicotiana tabacum . Plant J. 30, 625–637 [DOI] [PubMed] [Google Scholar]
  53. Hess W.R., Mueller A., Nagy F., Börner T. (1994). Ribosome-deficient plastids affect transcription of light-induced nuclear genes: genetic evidence for a plastid-derived signal. Mol. Gen. Genet. 242, 305–312 [DOI] [PubMed] [Google Scholar]
  54. Hirose T., Sugiura M. (2004). Multiple elements required for translation of plastid atpB mRNA lacking the Shine–Dalgarno sequence. Nucleic Acids Res. 32, 3503–3510 [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Hirose T., Kusumegi T., Sugiura M. (1998). Translation of tobacco chloroplast rps14 mRNA depends on a Shine–Dalgarno-like sequence in the 5’-untranslated region but not on internal RNA editing in the coding region. FEBS Lett. 430, 257–260 [DOI] [PubMed] [Google Scholar]
  56. Hopkins W.G., Elfman B. (1984). Temperature-induced chloroplast ribosome deficiency in virescent maize. J. Hered. 75, 207–211 [Google Scholar]
  57. Horiguchi G., Van Lijsebettens M., Candela H., Micol J.L., Tsukaya H. (2012). Ribosomes and translation in plant developmental control. Plant Sci. 191–192, 24–34 [DOI] [PubMed] [Google Scholar]
  58. Hricová A., Quesada V., Micol J.L. (2006). The SCABRA3 nuclear gene encodes the plastid RpoTp RNA polymerase, which is required for chloroplast biogenesis and mesophyll cell proliferation in Arabidopsis . Plant Physiol. 141, 942–956 [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Ingolia N.T., Brar G.A., Rouskin S., McGeachy A.M., Weissman J.S. (2012). The ribosome profiling strategy for monitoring translation in vivo by deep sequencing of ribosome-protected mRNA fragments. Nature Protoc. 7, 1534–1550 [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Ingolia N.T., Ghaemmaghami S., Newman J.R.S., Weissman J.S. (2009). Genome-wide analysis in vivo of translation with nucleotide resolution using ribosome profiling. Science. 324, 218–223 [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Jarvis P., López-Juéz E. (2013). Biogenesis and homeostasis of chloroplast and other plastids. Nat. Rev. Mol. Cell Biol. 14, 787–802 [DOI] [PubMed] [Google Scholar]
  62. Kahlau S., Bock R. (2008). Plastid transcriptomics and translatomics of tomato fruit development and chloroplast-to-chromoplast differentiation: chromoplast gene expression largely serves the production of a single protein. Plant Cell. 20, 856–874 [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Karcher D., Bock R. (2009). Identification of the chloroplast adenosine-to-inosine tRNA editing enzyme. RNA. 15, 1251–1257 [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Kawade K., Horiguchi G., Tsukaya H. (2010). Non-cell-autonomously coordinated organ size regulation in leaf development. Development. 137, 4221–4227 [DOI] [PubMed] [Google Scholar]
  65. Kim G.-T., Cho K.-H. (2006). Recent advances in the genetic regulation of the shape of simple leaves. Physiol. Plant. 126, 494–502 [Google Scholar]
  66. Kode V., Mudd E.A., Iamtham S., Day A. (2005). The tobacco plastid accD gene is essential and is required for leaf development. Plant J. 44, 237–244 [DOI] [PubMed] [Google Scholar]
  67. Koussevitzky S., Nott A., Mockler T.C., Hong F., Sachetto-Martins G., Surpin M., Lim I.J., Mittler R., Chory J. (2007). Signals from chloroplasts converge to regulate nuclear gene expression. Science. 316, 715–719 [PubMed] [Google Scholar]
  68. Krause K. (2011). Piecing together the puzzle of parasitic plant plastome evolution. Planta. 234, 647–656 [DOI] [PubMed] [Google Scholar]
  69. Krech K., Ruf S., Masduki F.F., Thiele W., Bednarczyk D., Albus C.A., Tiller N., Hasse C., Schöttler M.A., Bock R. (2012). The plastid genome-encoded Ycf4 protein functions as a nonessential assembly factor for photosystem I in higher plants. Plant Physiol. 159, 579–591 [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Lee K.-H., Park J., Williams D.S., Xiong Y., Hwang I., Kang B.-H. (2013). Defective chloroplast development inhibits maintenance of normal levels of abscisic acid in a mutant of the Arabidopsis RH3 DEAD-box protein during early post-germination growth. Plant J. 73, 720–732 [DOI] [PubMed] [Google Scholar]
  71. Leister D. (2012). Retrograde signaling in plants: from simple to complex scenarios. Frontiers in Plant Science. 3, 135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Lepistö A., Toivola J., Nikkanen L., Rintamäki E. (2012). Retrograde signaling from functionally heterogeneous plastids. Frontiers in Plant Science. 3, 286. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Liere K., Börner T. (2007). Transcription and transcriptional regulation in plastids. Top. Curr. Genet. 19, 121–174 [Google Scholar]
  74. Lin Q., Yu N. -J., Spremulli L.L. (1996). Expression and functional analysis of Euglena gracilis chloroplast initiation factor 3. Plant Mol. Biol. 32, 937–945 [DOI] [PubMed] [Google Scholar]
  75. Ma Z., Dooner H.K. (2004). A mutation in the nuclear-encoded plastid ribosomal protein S9 leads to early embryo lethality in maize. Plant J. 37, 92–103 [DOI] [PubMed] [Google Scholar]
  76. Majeran W., Wollman F.-A., Vallon O. (2000). Evidence for a role of ClpP in the degradation of the chloroplast cytochrome b6f complex. Plant Cell. 12, 137–149 [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Manuell A.L., Quispe J., Mayfield S.P. (2007). Structure of the chloroplast ribosome: novel domains for translation regulation. PLoS Biol. 5, e209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Meurer J., Grevelding C., Westhoff P., Reiss B. (1998a). The PAC protein affects the maturation of specific chloroplast mRNAs in Arabidopsis thaliana . Mol. Gen. Genet. 258, 342–351 [DOI] [PubMed] [Google Scholar]
  79. Meurer J., Lezhneva L., Amann K., Gödel M., Bezhani S., Sherameti I., Oelmüller R. (2002). A peptide chain release factor 2 affects the stability of UGA-containing transcripts in Arabidopsis chloroplasts. Plant Cell. 14, 3255–3269 [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Meurer J., Plücken H., Kowallik K.V., Westhoff P. (1998b). A nuclear-encoded protein of prokaryotic origin is essential for the stability of photosystem II in Arabidopsis thaliana . EMBO J. 17, 5286–5297 [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Miura E., Kato Y., Matsushima R., Albrecht V., Laalami S., Sakamoto W. (2007). The balance between protein synthesis and degradation in chloroplasts determines leaf variegation in Arabidopsis yellow variegated mutants. Plant Cell. 19, 1313–1328 [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Morden C.W., Wolfe K.H., dePamphilis C.W., Palmer J.D. (1991). Plastid translation and transcription genes in a non-photosynthetic plant: intact, missing and pseudo genes. EMBO J. 10, 3281–3288 [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Morita-Yamamuro C., Tsutsui T., Tanaka A., Yamaguchi J. (2004). Knock-out of the plastid ribosomal protein S21 causes impaired photosynthesis and sugar-response during germination and seedling development in Arabidopsis thaliana . Plant Cell Physiol. 45, 781–788 [DOI] [PubMed] [Google Scholar]
  84. Moschopoulos A., Derbyshire P., Byrne M.E. (2012). The Arabidopsis organelle-localized glycyl-tRNA synthetase encoded by EMBRYO DEFECTIVE DEVELOPMENT1 is required for organ patterning. J. Exp. Bot. 63, 5233–5243 [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Motohashi R., Yamazaki T., Myouga F., Ito T., Ito K., Satou M., Kobayashi M., Nagata N., Yoshida S., Nagashima A., et al. (2007). Chloroplast ribosome release factor 1 (AtcpRF1) is essential for chloroplast development. Plant Mol. Biol. 64, 481–497 [DOI] [PubMed] [Google Scholar]
  86. Mullet J.E., Klein R.R. (1987). Transcription and RNA stability are important determinants of higher plant chloroplast RNA levels. EMBO J. 6, 1571–1579 [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Naested H., Holm A., Jenkins T., Nielsen H.B., Harris C.A., Beale M.H., Andersen M., Mant A., Scheller H., Camara B., et al. (2004). Arabidopsis VARIEGATED 3 encodes a chloroplast-targeted, zinc-finger protein required for chloroplast and palisade cell development. J. Cell Sci. 117, 4807–4818 [DOI] [PubMed] [Google Scholar]
  88. Nishimura K., Ashida H., Ogawa T., Yokota A. (2010). A DEAD box protein is required for formation of a hidden break in Arabidopsis chloroplast 23S rRNA. Plant J. 63, 766–777 [DOI] [PubMed] [Google Scholar]
  89. Peled-Zehavi H., Danon A. (2007). Translation and translational regulation in chloroplasts. Top. Curr. Genet. 19, 249–281 [Google Scholar]
  90. Pesaresi P., Varotto C., Meurer J., Jahns P., Salamini F., Leister D. (2001). Knock-out of the plastid ribosomal protein L11 in Arabidopsis: effects on mRNA translation and photosynthesis. Plant J. 27, 179–189 [DOI] [PubMed] [Google Scholar]
  91. Pinon V., Etchells J.P., Rossignol P., Collier S.A., Arroyo J.M., Martienssen R.A., Byrne M.E. (2008). Three PIGGYBACK genes that specifically influence leaf patterning encode ribosomal proteins. Development. 135, 1315–1324 [DOI] [PubMed] [Google Scholar]
  92. Pyke K., Zubko M.K., Day A. (2000). Marking cell layers with spectinomycin provides a new tool for monitoring cell fate during leaf development. J. Exp. Bot. 51, 1713–1720 [DOI] [PubMed] [Google Scholar]
  93. Rauwolf U., Golczyk H., Meurer J., Herrmann R.G., Greiner S. (2008). Molecular marker systems for Oenothera genetics. Genetics. 180, 1289–1306 [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Reiter R.S., Coomber S.A., Bourett T.M., Bartley G.E., Scolnik P.A. (1994). Control of leaf and chloroplast development by the Arabidopsis gene pale cress. Plant Cell. 6, 1253–1264 [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Rhoades M.M. (1955). Interaction of genic and non-genic hereditary units and the physiology of non-genic inheritance. In Encyclopedia of Plant Physiology, Vol. 1, Ruhland W, ed. (Göttingen, Heidelberg: Springer Verlag Berlin; ), pp. 19–57 [Google Scholar]
  96. Rodermel S. (2001). Pathways of plastid-to-nucleus signaling. Trends Plant Sci. 6, 471–477 [DOI] [PubMed] [Google Scholar]
  97. Rogalski M., Karcher D., Bock R. (2008a). Superwobbling facilitates translation with reduced tRNA sets. Nat. Struct. Mol. Biol. 15, 192–198 [DOI] [PubMed] [Google Scholar]
  98. Rogalski M., Ruf S., Bock R. (2006). Tobacco plastid ribosomal protein S18 is essential for cell survival. Nucleic Acids Res. 34, 4537–4545 [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Rogalski M., Schöttler M.A., Thiele W., Schulze W.X., Bock R. (2008b). Rpl33, a nonessential plastid-encoded ribosomal protein in tobacco, is required under cold stress conditions. Plant Cell. 20, 2221–2237 [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Romani I., Tadini L., Rossi F., Masiero S., Pribil M., Jahns P., Kater M., Leister D., Pesaresi P. (2012). Versatile roles of Arabidopsis plastid ribosomal proteins in plant growth and development. Plant J. 72, 922–934 [DOI] [PubMed] [Google Scholar]
  101. Rosner A., Porath D., Gressel J. (1974). The distribution of plastid ribosomes and the integrity of plastid ribosomal RNA during the greening and maturation of Spirodela fronds. Plant Cell Physiol. 15, 891–902 [Google Scholar]
  102. Rott M., Martins N.F., Thiele W., Lein W., Bock R., Kramer D.M., Schöttler M.A. (2011). ATP synthase repression in tobacco restricts photosynthetic electron transport, CO2 assimilation, and plant growth by overacidification of the thylakoid lumen. Plant Cell. 23, 304–321 [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Ruf M., Kössel H. (1988). Occurrence and spacing of ribosome recognition sites in mRNAs of chloroplasts from higher plants. FEBS Lett. 240, 41–44 [Google Scholar]
  104. Ruppel N.J., Hangarter R.P. (2007). Mutations in a plastid-localized elongation factor G alter early stages of plastid development in Arabidopsis thaliana . BMC Plant Biol. 7, 37. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Scharff L.B., Bock R. (2014). Synthetic biology in plastids. Plant J., in press. [DOI] [PubMed] [Google Scholar]
  106. Scharff L.B., Childs L., Walther D., Bock R. (2011). Local absence of secondary structure permits translation of mRNAs that lack ribosome-binding sites. PLoS Genet. 7, e1002155. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Schmitz-Linneweber C., Small I. (2008). Pentatricopeptide repeat proteins: a socket set for organelle gene expression. Trends Plant Sci. 13, 663–670 [DOI] [PubMed] [Google Scholar]
  108. Schultes N.P., Sawers R.J.H., Brutnell T.P., Krueger R.W. (2000). Maize high chlorophyll fluorescent 60 mutation is caused by an Ac disruption of the gene encoding the chloroplast ribosomal small subunit protein 17. Plant J. 21, 317–327 [DOI] [PubMed] [Google Scholar]
  109. Schwarz Z., Kössel H. (1979). Sequencing of the 3’-terminal region of a 16S rRNA gene from Zea mays chloroplast reveals homology with E. coli 16S rRNA. Nature. 279, 520–522 [DOI] [PubMed] [Google Scholar]
  110. Schwarz Z., Kössel H. (1980). The primary structure of 16S rDNA from Zea mays chloroplast is homologous to E. coli 16S rRNA. Nature. 283, 739–742 [Google Scholar]
  111. Schwemmle J. (1938). Genetische und zytologische Untersuchungen an Eu-Oenotheren. III. Untersuchungen über das Zusammenwirken von Kern, Plasma und Plastiden. Z. indukt. Abst. u. Vererb.-Lehre. 75, 486–660 [Google Scholar]
  112. Schwemmle J. (1941). Weitere Untersuchungen an Eu-Oenotheren über die genetische Bedeutung des Plasmas und der Plastiden. Z. indukt. Abstammungs- u. Vererbungslehre. 79, 321–333 [Google Scholar]
  113. Schwemmle J. (1943). Plastiden und Genmanifestation. Flora. 137, 61–72 [Google Scholar]
  114. Sharma M.R., Dönhöfer A., Barat C., Marquez V., Datta P.P., Fucini P., Wilson D.N., Agrawal R.K. (2010). PSRP1 is not a ribosomal protein, but a ribosome-binding factor that is recycled by the ribosome-recycling factor (RRF) and elongation factor G (EF-G). J. Biol. Chem. 285, 4006–4014 [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Sharma M.R., Wilson D.N., Datta P.P., Barat C., Schluenzen F., Fucini P., Agrawal R.K. (2007). Cryo-EM study of the spinach chloroplast ribosome reveals the structural and functional roles of plastid-specific ribosomal proteins. Proc. Natl Acad. Sci. U S A. 104, 19315–19320 [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. Shen Y., Li C., McCarty D.R., Meeley R., Tan B.-C. (2013). Embryo defective12 encodes the plastid initiation factor 3 and is essential for embryogenesis in maize. Plant J. 74, 792–804 [DOI] [PubMed] [Google Scholar]
  117. Shikanai T., Shimizu K., Ueda K., Nishimura Y., Kuroiwa T., Hashimoto T. (2001). The chloroplast cplP gene, encoding a proteolytic subunit of ATP-dependent protease, is indispensable for chloroplast development in tobacco. Plant Cell Physiol. 42, 264–273 [DOI] [PubMed] [Google Scholar]
  118. Shimada H., Sugiura M. (1991). Fine structural features of the chloroplast genome: comparison of the sequenced chloroplast genomes. Nucleic Acids Res. 19, 983–995 [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Shoji S., Dambacher C.M., Shajani Z., Williamson J.R., Schultz P.G. (2011). Systematic chromosomal deletion of bacterial ribosomal protein genes. J. Mol. Biol. 413, 751–761 [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Staub J.M., Maliga P. (1993). Accumulation of D1 polypeptide in tobacco plastids is regulated via the untranslated region of the psbA mRNA. EMBO J. 12, 601–606 [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Stern D.B., Goldschmidt-Clermont M., Hanson M.R. (2010). Chloroplast RNA metabolism. Annu. Rev. Plant Biol. 61, 125–155 [DOI] [PubMed] [Google Scholar]
  122. Stoppel R., Meurer J. (2012). The cutting crew: ribonucleases are key players in the control of plastid gene expression. J. Exp. Bot. 63, 1663–1673 [DOI] [PubMed] [Google Scholar]
  123. Streatfield S.J., Weber A., Kinsman E.A., Häusler R.E., Li J., Post-Beittenmiller D., Kaiser W.M., Pyke K.A., Flügge U. -I., Chory J. (1999). The phophoenolpyruvate/phosphate translocator is required for phenolic metabolism, palisade cell development, and plastid-dependent nuclear gene expression. Plant Cell. 11, 1609–1621 [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. Sugiura M., Wakasugi T. (1989). Compilation and comparison of transfer RNA genes from tobacco chloroplasts. Crit. Rev. Plant Sci. 8, 89–101 [Google Scholar]
  125. Sugiura M., Hirose T., Sugita M. (1998). Evolution and mechanism of translation in chloroplasts. Annu. Rev. Genet. 32, 437–459 [DOI] [PubMed] [Google Scholar]
  126. Sullivan J.A., Gray J.C. (1999). Plastid translation is required for the expression of nuclear photosynthesis genes in the dark and in roots of the pea lip1 mutant. Plant Cell. 11, 901–910 [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Susek R.E., Ausubel F.M., Chory J. (1993). Signal transduction mutants of Arabidopsis uncouple nuclear cab and rbcS gene expression from chloroplast development. Cell. 74, 787–799 [DOI] [PubMed] [Google Scholar]
  128. Swiatecka-Hagenbruch M., Emanuel C., Hedtke B., Liere K., Börner T. (2008). Impaired function of the phage-type RNA polymerase RpoTp in transcription of chloroplast genes is compensated by a second phage-type RNA polymerase. Nucleic Acids Res. 36, 785–792 [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Sylvester A.W., Smith L., Freeling M. (1996). Acquisition of identity in the developing leaf. Annu. Rev. Cell Dev. Biol. 12, 257–304 [DOI] [PubMed] [Google Scholar]
  130. Tameshige T., Fujita H., Watanabe K., Toyokura K., Kondo M., Tatematsu K., Matsumoto N., Tsugeki R., Kawaguchi M., Nishimura M., et al. (2013). Pattern dynamics in adaxial-abaxial specific gene expression are modulated by a plastid retrograde signal during Arabidopsis thaliana leaf development. PLoS Genet. 9, e1003655. [DOI] [PMC free article] [PubMed] [Google Scholar]
  131. Tan W., Bögre L., López-Juez E. (2008). Light fluence rate and chloroplasts are sources of signals controlling mesophyll cell morphogenesis and division. Cell Biol. Int. 32, 563–565 [DOI] [PubMed] [Google Scholar]
  132. Tiller N., Weingartner M., Thiele W., Maximova E., Schöttler M.A., Bock R. (2012). The plastid-specific ribosomal proteins of Arabidopsis thaliana can be divided into non-essential proteins and genuine ribosomal proteins. Plant J. 69, 302–316 [DOI] [PubMed] [Google Scholar]
  133. Timmis J.N., Ayliffe M.A., Huang C.Y., Martin W. (2004). Endosymbiotic gene transfer: organelle genomes forge eukaryotic chromosomes. Nature Rev. Genet. 5, 123–136 [DOI] [PubMed] [Google Scholar]
  134. Tsukaya H. (1995). Developmental genetics of leaf morphogenesis in dicotyledonous plants. J. Plant Res. 108, 407–416 [Google Scholar]
  135. Ueda M., Fujimoto M., Arimura S.-i., Murata J., Tsutsumi N., Kadowaki K.-i. (2007). Loss of rpl32 gene from the chloroplast genome and subsequent acquisition of a preexisting transit peptide within the nuclear gene in Populus. Gene. 402, 51–56 [DOI] [PubMed] [Google Scholar]
  136. Valkov V.T., Scotti N., Kahlau S., MacLean D., Grillo S., Gray J.C., Bock R., Cardi T. (2009). Genome-wide analysis of plastid gene expression in potato leaf chloroplasts and tuber amyloplasts: transcriptional and posttranscriptional control. Plant Physiol. 150, 2030–2044 [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Vogel J., Hübschmann T., Börner T., Hess W.R. (1997). Splicing and intron-internal RNA editing of trnK-matK transcripts in barley plastids: support for matK as an essential splicing factor. J. Mol. Biol. 270, 179–187 [DOI] [PubMed] [Google Scholar]
  138. Wakasugi T., Tsudzuki T., Sugiura M. (2001). The genomics of land plant chloroplasts: gene content and alteration of genomic information by RNA editing. Photosynthesis Res. 70, 107–118 [DOI] [PubMed] [Google Scholar]
  139. Walter M., Piepenburg K., Schöttler M.A., Petersen K., Kahlau S., Tiller N., Drechsel O., Weingartner M., Kudla J., Bock R. (2010). Knockout of the plastid RNase E leads to defective RNA processing and chloroplast ribosome deficiency. Plant J. 64, 851–863 [DOI] [PubMed] [Google Scholar]
  140. Wang L., Ouyang M., Li Q., Zou M., Guo J., Ma J., Lu C., Zhang L. (2010). The Arabidopsis chloroplast ribosome recycling factor is essential for embryogenesis and chloroplast biogenesis. Plant Mol. Biol. 74, 47–59 [DOI] [PubMed] [Google Scholar]
  141. Wang Y., Duby G., Purnelle B., Boutry M. (2000). Tobacco VDL gene encodes a plastid DEAD box RNA helicase and is involved in chloroplast differentiation and plant morphogenesis. Plant Cell. 12, 2129–2142 [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Wanke D. (2011). The ABA-mediated switch between submersed and emersed life-styles in aquatic macrophytes. J. Plant Res. 124, 467–475 [DOI] [PubMed] [Google Scholar]
  143. Wycliffe P., Sitbon F., Wernersson J., Ezcurra I., Ellerström M., Rask L. (2005). Continuous expression in tobacco leaves of a Brassica napus PEND homologue blocks differentiation of plastids and development of palisade cells. Plant J. 44, 1–15 [DOI] [PubMed] [Google Scholar]
  144. Xu T., Lee K., Gu L., Kim J.-I., Kang H. (2013). Functional characterization of a plastid-specific ribosomal protein PSRP2 in Arabidopsis thaliana under abiotic stress conditions. Plant Physiol. Biochem. 73, 405–411 [DOI] [PubMed] [Google Scholar]
  145. Yamaguchi K., Subramanian A.R. (2000). The plastid ribosomal proteins: identification of all the proteins in the 50 S subunit of an organelle ribosome (chloroplast). J. Biol. Chem. 275, 28466–28482 [DOI] [PubMed] [Google Scholar]
  146. Yamaguchi K., Subramanian A.R. (2003). Proteomic identification of all plastid-specific ribosomal proteins in higher plant chloroplast 30S ribosomal subunit. Eur. J. Biochem., 190–205 [DOI] [PubMed] [Google Scholar]
  147. Yamaguchi K., von Knoblauch K., Subramanian A.R. (2000). The plastid ribosomal proteins: identification of all the proteins in the 30 S subunit of an organelle ribosome (chloroplast). J. Biol. Chem. 275, 28455–28465 [DOI] [PubMed] [Google Scholar]
  148. Yao Y., Ling Q., Wang H., Huang H. (2008). Ribosomal proteins promote leaf adaxial identity. Development. 135, 1325–1334 [DOI] [PubMed] [Google Scholar]
  149. Yin T., Pan G., Liu H., Wu J., Li Y., Zhao Z., Fu T., Zhou Y. (2012). The chloroplast ribosomal protein L21 gene is essential for plastid development and embryogenesis in Arabidopsis . Planta. 235, 907–921 [DOI] [PubMed] [Google Scholar]
  150. Zoschke R., Watkins K.P., Barkan A. (2013). A rapid ribosome profiling method elucidates chloroplast ribosome behavior in vivo . Plant Cell. 25, 2265–2275 [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Molecular Plant are provided here courtesy of Oxford University Press

RESOURCES