Next Article in Journal
Antioxidant Activity Improvement of Apples Juice Supplemented with Chitosan-Galactose Maillard Reaction Products
Previous Article in Journal
Degradation of Tocopherol Molecules and Its Impact on the Polymerization of Triacylglycerols during Heat Treatment of Oil
Previous Article in Special Issue
Biosynthesis of Isoprene Units in Euphorbia lathyris Laticifers vs. Other Tissues: MVA and MEP Pathways, Compartmentation and Putative Endophytic Fungi Contribution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Functional Gene Network of Prenyltransferases in Arabidopsis thaliana

1
Institute of Biology and Ecology, Pavol Jozef Šafárik University in Košice, 04180 Košice, Slovakia
2
Institute of Pharmacology, University of Veterinary Medicine and Pharmacy in Košice, 04181 Košice, Slovakia
*
Author to whom correspondence should be addressed.
Molecules 2019, 24(24), 4556; https://doi.org/10.3390/molecules24244556
Submission received: 7 November 2019 / Revised: 9 December 2019 / Accepted: 10 December 2019 / Published: 12 December 2019
(This article belongs to the Special Issue Plant Isoprenoids)

Abstract

:
Prenyltransferases (PTs) are enzymes that catalyze prenyl chain elongation. Some are highly similar to each other at the amino acid level. Therefore, it is difficult to assign their function based solely on their sequence homology to functional orthologs. Other experiments, such as in vitro enzymatic assay, mutant analysis, and mutant complementation are necessary to assign their precise function. Moreover, subcellular localization can also influence the functionality of the enzymes within the pathway network, because different isoprenoid end products are synthesized in the cytosol, mitochondria, or plastids from prenyl diphosphate (prenyl-PP) substrates. In addition to in vivo functional experiments, in silico approaches, such as co-expression analysis, can provide information about the topology of PTs within the isoprenoid pathway network. There has been huge progress in the last few years in the characterization of individual Arabidopsis PTs, resulting in better understanding of their function and their topology within the isoprenoid pathway. Here, we summarize these findings and present the updated topological model of PTs in the Arabidopsis thaliana isoprenoid pathway.

1. Introduction

Isoprenoids form a large group of more than 55,000 natural metabolites despite the fact that they are all derived from just five carbon units of isopentenyl diphosphate (IPP) [1]. IPP and its isomer dimethylallyl diphosphate (DMAPP) are synthetized in a highly conserved manner in almost all living species. In plants, they are produced in the cytosol via the mevalonic acid (MVA) pathway and in plastids through the MVA-independent methylerythritol phosphate (MEP) pathway. However, the vast majority of organisms possess just one of these pathways—the MVA pathway is more widely spread, utilized by archaebacteria, some Gram-positive bacteria, yeasts, fungi, and animals, and the MEP pathway provides the isoprenoid precursors in bacteria, cyanobacteria, and green algae [2].
Some groups of isoprenoids, such as cytokinins and hemiterpenoids, are derived directly from the DMAPP, but the great structural and functional diversity of isoprenoids is created by the activity of prenyltransferases (PTs). In a broader sense, PTs are any enzymes that catalyze the transfer of the prenyl group [3]. Possible acceptors are both aliphatic and aromatic compounds with low or even high molecular weight including proteins. In a narrower sense, PTs catalyze condensation between DMAPP or another prenyl-PP, which serves as an allylic donor, and IPP, which serves as an acceptor [4] (Figure 1). These are also known in the literature as isoprenyl-PP synthases (IDSs) [5]. In this review, we will use the term PT for convenience, as PT is more frequently used in the literature, but we will refer to only the prenyl chain elongating enzymes that catalyze sequential condensation of allylic diphosphate and IPP.
The binding of both co-substrates to the enzyme in the presence of the divalent Mg salts creates carbocation on the allylic substrate, followed by the electrophilic attack on the double bond of the IPP. As a result, a new carbon bond joining the co-substrates is formed [6]. The product is then released from the active site or can act as a primer in the subsequent coupling with another IPP molecule. With every further elongation step, a linear prenyl-PP grows by C5. The carbon chain length of the final isoprenoid polymer and its stereochemical properties are determined by the specificity of the PT, which is given to the PT by its structural features. The two main classes of PTs are recognized due to the differences in the spatial arrangement of units in the isoprenoid product, namely trans- or (E)-PTs and cis- or (Z)-PTs (Figure 1). Basically, shorter prenyl-PPs are formed via trans-condensation and polyprenyl-PPs via cis-condensation in plants [7]. Although trans- and cis-PTs catalyze similar reactions, they are evolutionarily distinct [7] (Figure 2).

2. Prenyltransferases in Arabidopsis thaliana

IPP is a prenyl acceptor of the prenyl group originating from DMAPP, in the reaction catalyzed by PTs. IPP and DMAPP, both substrates of PTs, are interconvertible isomers, whose equilibrium is controlled by IPP isomerase (IPPI) [12]. The enzyme is encoded by IPPI1 (At3g02780) and IPPI2 (At5g16440) genes in A. thaliana. Both are transcribed into two different protein variants—short with peroxisomal targeting (IPPI1S, IPPI2S), supplying IPP for the cytosolic PTs, and long with either mitochondrial (IPPI1L) or plastidial targeting (IPPI2L), providing IPP for PTs in respective organelles [13,14,15] (Figure 3). In agreement with that, PT activity at the level of a plant cell is also compartmentalized [16].

2.1. Trans-Prenyltransferases

C–C double bonds in prenyl-PP can be either in trans or cis conformation depending on the PT—trans-PTs form trans isomers, and cis-PTs form cis-isomers. Considering the length of the produced carbon backbone, trans-PTs can be classified as short-chain PTs (C10–25), medium-chain PTs (C30–35), and long-chain PTs (C40–50) [7]. Members of the short-chain ones are in ascending order C10 geranyl diphosphate synthase (GPPS), C15 farnesyl diphosphate synthase (FPPS), C20 geranylgeranyl diphosphate synthase (GGPPS), and C25 geranylfarnesyl diphosphate synthase (GFPPS) (Table 1). Their respective products serve as substrates for several downstream enzymes and, as such, represent important branch points in isoprenoid biosynthesis.
C10 geranyl diphosphate (GPP), the precursor of monoterpenoids, is the simplest product of the condensation reaction of DMAPP with IPP that is controlled by GPPS residing in plastids [8,17] (Figure 3, Table S1). Initially, the GPPS function was attributed to a homodimeric protein (At2g34630) [18]. Later, however, it was revealed that GPPS in A. thaliana exists only as a heterodimer composed of one large subunit (LSU) and one small subunit (SSU) [8,9]. While LSU as a homodimer also functions as GGPPS (GGPPS11, At4g36810), its activity is shifted towards enhanced GPP production, when it dimerazes with SSU (At4g38460). SSU itself is not catalytically active due to the lack of one of the two aspartate-rich motifs, which were shown to play a role in binding prenyl-PP substrates by PTs [19,20] (Figure 4).
For the physical interaction between the subunits, the CXXXC (X is any hydrophobic amino acid residue) motif was found to be critical [9]. This motif is characteristic particularly for the SSU subfamily, but it is also present in GGPP and GFPP synthases and in polyprenyl-PP synthase 2 (PPPS2) (Figure 4). In A. thaliana, this type of interaction was reported specifically for GGPPS11 and SSU, earlier annotated as GGPPS12 [24]. Because both enzymes reside in plastids, GPP synthesis can be assigned solely to plastids. This is in agreement with monoterpenoid biosynthesis, which is located in plastids [16].
FPP is formed from DMAPP or GPP and IPP via action of FPPS [25] (Table S1). It is a major substrate for cytosolic and mitochondrial branches of isoprenoid biosynthesis (Figure 3). FPP-derived compounds, such as sterols, brassinosteroids, dolichols, and ubiquinones (UQ), are essential for plant growth and development. Others, such as triterpenoids and sesquiterpenoids, mediate defense reactions. Last, but not least, FPP plays a very important role in protein prenylation and hence in cell signaling [26]. In A. thaliana, two FPPS genes were identified and characterized—FPPS1 (At5g47770) with two protein variants and FPPS2 (At4g17190) [25,27,28,29,30,31,32]. The longer-variant FPPS1L is targeted to mitochondria, the shorter FPPS1S, together with the FPPS2, is localized to the cytosol. Research on this topic demonstrated that FPPS genes are redundant, thus one copy of the gene is enough to keep plants viable with only small changes in metabolite content [25,30,31,32,33]. On the other hand, total loss of FPPS activity is lethal due to severely impaired embryo development [31].
GGPP is a product of GGPPS, which accepts allylic DMAPP, GPP, and FPP in the condensation reaction with IPP [5,9,23] (Table S1). In plant cells, GGPP is used for the synthesis of gibberellins and other diterpenoids, as well as carotenoids; it also participates in chlorophyll, tocopherol, plastoquinone (PQ), UQ, and phylloquinone biosynthesis and enables protein geranylgeranylation (Figure 3). GGPPS is encoded by a gene family in A. thaliana. At first, 12 gene paralogs were predicted to form the GGPPS gene family (GGPPS1-12) [24], but with progressive research in the field and more sophisticated methods of analysis, several gene paralogs were excluded from this family.
One of them was GGPPS5 (At3g14510). Because of a frame-shift mutation in its nucleotide sequence, GGPPS5 encodes a nonfunctional protein and is therefore considered a pseudogene [17]. SSU has no GGPPS activity on its own, but upon dimerization with GGPPS11, changes its product specificity more towards GPP production [9]. In recent years, all the remaining members of the GGPPS family were reexamined for their function, and the analysis brought up surprising results (Table S1). Two independent studies provided evidence that GGPPS6 (At3g14530), GGPPS7 (At3g14550), GGPPS9 (At3g29430), and GGPPS10 (At3g32040) encode GFPPS enzymes rather than GGPPS [5,23]. Their names were therefore changed to GFPPS1, GFPPS2, GFPPS3, and GFPPS4, respectively [23]. Additionally, GGPPS8 (At3g20160) also failed to synthesize GGPP [5,23] and instead produced a mix of prenyl-PPs ranging from C20 to C40, with the predominant C30–C35 species. GGPPS8 was therefore renamed to PPPS2 [23], and together with PPPS1 (At2g34630), represents the medium chain-length PTs in A. thaliana. The remaining GGPPS enzymes showed solely or predominantly GGPPS activity and therefore remained members of the A. thaliana GGPPS family. However, the GGPPS1 (At1g49530) classification is still controversial, because Nagel et. al. [5] showed that this enzyme has both GFPPS and GGPPS activity, with GFPPS activity being the prominent one, whereas Wang et al. [23] showed that this enzyme has only GGPPS activity (Table S1). Thus, GGPPS1 is considered to be a member of the GGPPS gene family, which has five members: mitochondrial GGPPS1, plastidial GGPPS2 (At2g18620), GGPPS11, endoplasmic reticulum (ER)-bound GGPPS3 (At2g18640), and GGPPS4 (At2g23800). Their activities were confirmed in several studies by in vitro enzymatic assays, by genetic complementation of GGPPS-deficient Escherichia coli, or by analysis of A. thaliana mutants [5,9,17,23,34,36,53,54].
C25 GFPP is a precursor of sesterterpenoids, which are compounds with multiple bioactive properties and thus promising pharmacological potential [55]. They were occasionally reported in plants, but their origin had not been addressed until recently. Only with the discovery of GFPPS activity in A. thaliana has the topic become more attractive. GFPP is in A. thaliana synthesized by GFPPS1 (At3g14530), GFPPS2 (At3g14550), GFPPS3 (At3g29430), and GFPPS4 (At3g32040) (Figure 3). In in vitro tests, these isozymes utilized preferentially GPP, FPP, and GGPP as prenyl donors and, with lower efficiency, DMAPP as well [5,23] (Table S1). Although GFPP is a predominant product of the GFPPSs, the enzymes also have considerable GPPS and GGPPS activity [5,23]. This can explain why complementation tests of carotogenic E. coli lacking functional GGPPS provided false positive results when using GFPPS [17,23]. GFPPS genes, at least in Brassicaceae, occur in the vicinity of the terpene synthase (TPS) genes forming PT-TPS clusters [23,56,57].
Long-chain trans-PTs in A. thaliana are represented by solanesyl diphosphate synthases (SPPSs). Their C45 long product solanesyl diphosphate (SPP) can be found in prenylquinones as a prenyl side chain, which serves to anchor the molecule into the membrane [58]. Plant prenylquinones operate in electron transfer during photosynthesis in plastids (PQ) and during respiration in mitochondria (UQ). The length of their prenyl attachments vary among species—in A. thaliana plastoquinone 9 (PQ-9) and ubiquinone 9 (UQ-9) are predominant prenylquinone species [45], which means, they have mainly nonaprenyls (C45) attached. The first Arabidopsis SPPS1 (At1g78510) gene was described in 2003 by Hirooka et al. [59] The sequence was identified based on the similarity with Arabidopsis FPP and GGPP synthases and also bacterial long-chain PTs. According to in vitro enzymatic reaction, SPPS1 could utilize FPP and GGPP but not DMAPP and GPP as substrates. Its predominant product was SPP (Table S1). Another long-chain PT (SPPS2; At1g17050), with 79% amino acid sequence identity to SPPS1 (At1g78510), was described by Hirooka et al. in 2005 [46]. SPPS2, similar to SPPS1, used only FPP or GGPP but not DMAPP with IPP to synthesize SPP (Table S1). In vivo activity of both enzymes was checked in the background of Schizosaccharomyces pombe dlp1 and dps1 mutants lacking functional decaprenyl-PP synthase [45]. When either of the mutants expressed SPPS1 or SPPS2, its ability to grow on minimal medium was restored. Both enzymes reside in plastids [42,44,45,46] (Figure 3), although SPPS1 was initially described to localize to the ER [45,46]. In plastids, both enzymes interact with fibrillin FBN5, which is believed to be an indispensable structural component of PQ-9 biosynthesis through the direct stimulation of SPPS activity [44]. On the other hand, FBN5 did not interact with a plastidial fraction of PPPS1 (PPPS1pl; At2g34630), which is a medium-chain polyprenyl-PP synthase also able to catalyze the solanesyl formation [41,43]. PPPS1 was firstly characterized as a homomeric GPPS, active in plastidial monoterpenoid biosynthesis [18]. The corresponding gene shared 24.2%–28.7% identity with Arabidopsis short-chain PTs known at that time. When Hirooka et al. studied SPPS1 [59] and SPPS2 [46], they noticed a relatively high level of sequence identity between their genes and the previously misidentified GPPS (39% and 40%, respectively). However, the information was communicated very marginally, and only later did it draw the needed attention. Hsieh et al. [41] found out that putative GPPS can make C20–C45 products out of GPP, FPP, or GGPP when IPP is in excess (Table S1). They reasoned that previous in vitro testing of GPPS could have been negatively affected by inappropriate reaction conditions in terms of the used substrates and their mutual ratio. Finally, the enzyme’s participation in UQ biosynthesis was proved by successful complementation of the yeast coq1 mutant [43]. PPPS1 has dual mitochondrial and plastidial targeting, probably thanks to a genuine bifunctional pre-sequence [60] (Figure 3). In PPPS1-silenced and overexpressing Arabidopsis lines only the amount of UQ-9 but not PQ-9 was found to vary [43]. Thus, PPPS1 was proposed to synthesize the side chain of UQ-9 in mitochondria but not that of PQ-9 in plastids. Summarizing the data, we can conclude that SPPS1 and SPPS2 contribute to PQ-9 biosynthesis, whereas PPPS1 is involved in UQ-9 biosynthesis. This functional difference is further supported by their separate evolutionary history—SPPS1 and SPPS2 are close relatives of cyanobacterial trans-long-chain PTs, whereas PPPS1 has strictly eukaryotic origin [42].

2.2. Cis-Prenyltransferases

In plants, cis-prenyltransferases (cis-PTs or CPTs) catalyze the synthesis of polyprenols and dolichols that are usually longer than C50 in length [61]. Because CPTs use short all-trans-prenyl diphosphates—mostly FPP and GGPP as priming molecules—which conjugate with IPP in a cis orientation (Table S1), the resulting products exhibit mixed di-trans-poly-cis or tri-trans-poly-cis stereoisomerism, respectively [62] (Figure 1). While polyprenols are unsaturated at the terminal α-unit and are common in bacteria and plant photosynthetic tissues, dolichols are α-saturated and typically occur in yeasts, animals, and plant roots [49]. In fact, most organisms accumulate both forms of these polyisoprenoid alcohols, but the composition and ratio vary from species to species and also in a tissue-specific manner. The dynamic nature of polyisoprenoid metabolism can be observed during development or in response to various environmental stimuli suggesting diverse biological functions of these compounds [47]. Polyprenols, as well as dolichols, have the capacity to modulate the properties of membranes, they can serve as donors for protein prenylation, and dolichols, in particular, are inevitable saccharide carriers in protein glycosylation. Moreover, polyprenols that are quite abundant in chloroplasts can affect photosynthetic efficiency [51]. Current knowledge on the polyisoprenoid synthesis and function in plant cells is rather restricted and regulatory mechanisms controlling their biosynthesis are far from being understood.
Searching for plant CPTs, nine yeast CPT homologues (CPT1-9) and gene-encoding CPT-binding protein (CBP) LEW1 were identified in the A. thaliana genome (Table 1). So far, only four CPTs have been characterized and proven to be functional CPT enzymes. The first Arabidopsis CPT (CPT1, At2g23410) was described by the independent studies of Oh et al. [63] and Cunillera et al. [64]. Interestingly, the protein produced a different spectrum of polyprenols in vitro and in planta (more than C100) [49,63] and in vivo using rer2Δ yeasts that lack intrinsic CPT activity (less than C100) [49,64]. Moreover, the range of detected polyisoprenoids differed slightly even between roots and leaves in Arabidopsis CPT1 overexpressing lines [49]. Species and tissue-specific differences in chain length that were observed may be the result of different regulation of polyisoprenoid biosynthesis in different species and tissue background or may result from the availability of different substrates. Biosynthesis of short-chain polyisoprenoids in A. thaliana is mediated by heptaprenyl-PP synthase CPT6/HEPS (At5g58780), which catalyzes the formation of dolichol with seven isoprenoid units (Dol-7) [48,50] (Table S1). Both CPT1 [49] and CPT6 [48,50] were found to reside in the ER (Figure 3). CPT7 (At5g58770) is so far the only plastidial CPT that was characterized in Arabidopsis [51] (Figure 3). Following its subcellular localization, GGPP not FPP is likely the substrate for the synthesis of C55 polyisoprenoids, and mutant analysis of the cpt7-1 null mutant showed that CPT7 influences photosynthetic performance via modulation of thylakoid membrane dynamics.
There are two types of CPTs in plants—homomeric and heteromeric [65]. Homomeric ones act autonomously in the contrary to the heteromeric CPT enzymes, which need to dimerize with a CBP to gain its activity [66,67,68]. So far, the only CBP identified in A. thaliana is LEW1 (At1g11755), which was described to form a heterocomplex with CPT3 (At2g17570). It was shown that CPT3/LEW1 synthesizes long-chain polyisoprenoids, whereas individual CPT3 or LEW1 proteins were catalytically inactive [10]. Synthesis of polyisoprenoids by CPT/CBP heterocomplexes is typical for the eukaryotic lineage of CPTs, based on the phylogenetic analysis. Besides CPT3 (At2g17570), CPT4 (At5g60510) and CPT5 (At5g60500) also belong to this lineage in A. thaliana [48]. The remaining Arabidopsis CPTs cluster together in a prokaryotic clade, suggesting that they are self-sufficient homomeric enzymes. Recently, it was shown, that CBPs and homomeric CPTs have in common a highly conserved RXG (eventually NXG) motif in their C-terminal sequence, which is essential for catalysis [65] (Figure S1). On the other hand, the motif is absent in the heteromeric CPTs, and therefore, interaction with a CBP is required. Furthermore, CBPs are not able to bind substrates due to deletion, and thus, creating a CPT/CBP dimer is essential for the function of both proteins.

3. Primary Structure of PTs Can Be Indicative for Determination of Their Function

Protein functions are closely related to their structures. However, the structure was determined only for a small fraction of the myriad protein sequences deposited to databases. Therefore, functional characterization of novel sequences still heavily relies on homology-based approaches, in which experimentally determined functions are transferred to uncharacterized homologous sequences with various levels of confidence [69]. Such methods use the fact that homologous sequences sharing a high level of identity (at least 40%) and similarity are often isofunctional. However, Arabidopsis GGPP and GFPP synthases are functionally different, although they share more than 55% sequence identity [23]. Thus, the sequence alignment is not robust enough to assign the function to a protein, although it is still a convenient method for rough classification of unknown proteins.
Five highly conserved regions (I–V) can be recognized in the sequence of every trans-PT [70] (Figure 4). Regions II and V include aspartate-rich motifs DDX(2-4)D, referred to as the first aspartate-rich motif (FARM) and the second aspartate-rich motif (SARM). These motifs and +5 and +6 arginines (RR) downstream from the FARM are essential for catalysis and binding of the allylic substrate [19,20]. The attachment of IPP is coordinated by the C terminus of the enzyme and dispersed conserved basal residues [22]. From the point of the tertiary structure, all conserved features are associated with the active site, which is located in a large deep cleft of the enzyme formed out solely of α-helices [23,41]. The capacity of the catalytic pocket, where prenyl-PP elongation takes place, is determined by residues, which protrude into the cavity. Residues with large side chains can obstruct the cavity and thus further prenyl elongation [71]. It is currently unknown whether the mechanism is also valid for CPTs.
To describe the residues, which are an obstacle in the enzyme elongation channel, several authors have adopted the term “floor”. So far, single-, double-, and three-floor models of trans-PT chain-length determination have been proposed [23,41,71,72,73,74]. The three-floor model suggested by Wang et al. [23] is the most precise one. The model was deduced from the observations made on the crystal structures of three representatives of Arabidopsis PTs—GGPPS11, GFPPS2, and PPPS2. The authors revealed that some residues within the enzyme pocket assemble to form a floor, and there are three such floors along the pocket and thus three potential points where the synthesis of a prenyl product can be arrested due to steric hindrance from the blocking residue(s). Following the model scheme, the FPPS activity can be predicted from the two bulky residues (Tyr and Phe) on floor 1 [71] (Figure 4).
The GGPPS activity could be predicted from Phe on floor 3 and medium-sized Met on floor 1, which only partially closes the cavity (Figure 4). If solely Phe on floor 3 is present, the reaction should terminate with the release of GFPP (Figure 4). Dual GGPP/GFPP function could be indicated by Arg on floor 2 and Met on floor 3, which were detected in GGPPS1 (Figure 4). Finally, PPPS activity could be deduced from the lack of blocking residues, because there is only Met on floor 3 in PPPS2 (Figure 4). Predictions using the three-floor model are in agreement with experimental data on the product specificity of selected enzymes; thus, the model is functional at least with Arabidopsis short- and medium-chain trans-PTs. However, it would be interesting to see if it also works with long-chain trans-PTs. Mapping the model’s sites on the respective residues in PPPS1 has shown that no bulky residue is present in the pocket, but there are two in the SPPS1 and SPPS2—two molecules of Phe on floor 1 and 2 (Figure 4). Following the concept of the three floors model, the lack of obstacles in PPPS1 suggests that it is indeed PPPS. On the other hand, two residues blocking the cavity in SPPS are rather unlikely and hence challenge the relevance of the model with respect to the prediction of long-chain trans-PTs. In that case, the crystal structure of SPPS would be needed to optimize the model, but so far it is not available. Yet, we can see that in case of PTs, not the total sequence similarity but either the presence of specific domains or amino acid residues have predictive power when analyzing the amino acid sequence.
In past decades, determination of the key residues forming the pocket had been preceded by resolving the three-dimensional (3D) structure of the enzyme and/or laborious site-directed mutagenesis experiments [22,23,41,73,74,75,76]. In recent years, however, we can witness considerable efforts to create in silico prediction models, which could reliably assign function to a protein based on its raw sequence [1,77,78]. Naturally, 3D conformations remain an invaluable source of data needed to design a powerful algorithm.

4. Prenyltransferase Mutants: A Source of Information about Their Function and Topology

Analysis of knockouts and knockdowns is the way to understand the functionality of pathway enzymes within an in planta context. Visible phenotypes, moreover, can have predictive function for the topology of yet uncharacterized enzymes and for our knowledge on carbon fluxes via the isoprenoid pathway network. Both specific inhibitors and gene mutants were extensively used to analyze early enzymatic steps in the isoprenoid pathway (MVA and MEP pathways). In the case of PTs, however, mainly mutants were used to characterize their in planta function (Table 2).
Mutants in the linear metabolic pathway are expected to have similar phenotypes. MVA pathway null mutants are embryo lethal or male sterile, affected in male gametophyte development or in pollen tube elongation, whereas MEP pathway mutants grow to the seedling stage, but remain albinos [2]. In fact, there is only one null pathway mutant in the MVA pathway that is embryo and not male gametophyte lethal. It is an aact2 mutant impaired in acetoacetyl-CoA transferase, acting at the beginning of the MVA pathway [79]. It is expected that mutants in essential downstream pathways stemming from either MVA or MEP pathway will have identical or similar phenotypes. Indeed, mutants in the cytosolic/mitochondrial essential pathways, such as the sterol biosynthetic pathway, UQ biosynthetic pathway, protein Rab geranylgeranylation pathway, and dolichol biosynthetic pathway, have either impaired male gametophyte function or are embryo lethal [80,81,82,83]. Similarly, mutants in essential biosynthetic plastidial pathways, such as the carotenoid biosynthetic pathway, chlorophyll biosynthetic pathway, and PQ biosynthetic pathway, are albino or pale green [36,42,84,85]. In agreement with that, mutants in PTs that provide substrates for those specialized isoprenoid pathways should display similar phenotypes.

4.1. Prenyltransferase Mutants in the Cytosolic/Mitochondrial Branch of the Isoprenoid Pathway

The first step in the DMAPP and IPP condensation in the cytosol and mitochondria is FPP synthesis (Figure 3). There are two enzymes (FPPS1S, FPPS2) in the cytosol and one enzyme (FPPS1L) in the mitochondria that synthesize FPP. FPPS1L is encoded by the same locus as FPPS1S but only has mitochondrial targeting sequence. Neither the fpps1 nor the fpps2 null mutant (the fpps1 mutant is mutant in both FPPS1S and FPPS1L enzymes) shows the visible phenotype, and the levels of sterols and UQ-9, the major cytosolic and mitochondrial isoprenoids, are only slightly reduced in single mutants (Table 2). This demonstrates that both enzymes are highly redundant in Arabidopsis. Moreover, it shows that the function of FPPS1L in the mitochondria is not essential, excluding FPPS1L from being the only source of substrate for UQ-9 biosynthesis.
The fpps1fpps2 double mutant is embryo lethal, arrested at the globular stage of development. The fpps1fpps2 male gametophyte is only slightly affected in pollen tube elongation, which means that compared with MVA pathway mutants, which are mostly male gametophyte lethal, the fpps1fpps2 phenotype is less severe [2,31]. FPP is in the cytosol mainly used for the synthesis of phytosterols and the null mutant in the cycloartenol synthase (cas1), which synthesizes cycloartenol for the phytosterol biosynthetic pathway, is also male gametophyte lethal, as most of the MVA pathway mutants [81]. The less severe phenotype of fpps1fpps2 as compared with other mutants in sterol biosynthesis indicates that FPPS activity can be partly complemented, at least in some developmental stages. So far, we can only speculate about the source of the complementation. FPP can be imported from the wild-type cells of surrounding tissues or be synthesized by other PTs. For example, GGPPS11, which is also present in the cytosol, can synthesize a small amount of FPP (2.1%) [9] (Table S1).
The second step in DMAPP and IPP condensation in the cytosol and mitochondria is GGPP synthesis (Figure 3). There are three GGPPS enzymes in the cytosol (GGPPS3, GGPPS4, and GGPPS11S) and one in the mitochondria (GGPPS1). GGPPS11 transcript is, among all the GGPPS paralogs, the most abundant and the only one ubiquitously expressed in plants [17]. So far, ggpps11 and ggpps1 mutants have been isolated and characterized (Table 2). ggpps1 has no visible phenotype and no decrease in metabolite levels, including UQ-9 [36]. On the contrary, the ggpps11 mutant showed embryo lethal phenotype with the embryo arrested at the heart stage of development [36,54]. A similar phenotype is typical for the mutants deficient in the UQ biosynthesis [80]. The level of UQ-9 was tested in the ggpps11-5 knockdown mutant, which has decreased GGPPS11 transcript levels to about 50%. In this mutant, the level of UQ-9 was slightly decreased but not significantly [36], suggesting that the lack of UQ-9 is not the (main) reason for the embryo lethal phenotype of ggpps11 and suggesting, together with the results of ggpps1 analysis, that FPP rather than GGPP could be the source of prenyl-PP substrate for the UQ biosynthesis.
The level of UQ-9 was decreased significantly in fpps1 and fpps2 single mutants [31] and FPPS-silenced plants [32]. However, approximately 50% decrease in total FPPS activity, at the 16% level of transcripts for FPPS1 and 35% for FPPS2 in FPPS-silenced plants, caused only 30% decrease in UQ-9 levels [32]. This suggests that although FPPS1 and FPPS2 are contributing significantly to UQ biosynthesis, they are not the only substrates for SPP and subsequently UQ-9 biosynthesis. C45 solanesyl moiety, the predominant form of UQ prenyl chains in Arabidopsis, is synthesized by a mitochondrial PPPS1 (PPPS1mt) that elongates an initial FPP or GGPP with IPP units [41,43]. The ppps1 null mutant is, similarly to other UQ biosynthetic mutants, embryo lethal [80,87]. The reduction of PPPS1 transcript levels by 80% in PPPS1-silenced plants caused almost proportional reduction in UQ-9 content (reduction by 75–80%) [43]. This indicates that while PPPS1 is the only enzyme synthesizing C45 for UQ-9 biosynthesis, FPPS and GGPPS will have a rather promiscuous function in this process.
The question of whether the lack of GGPPS activity in the cytosol and mitochondria can to some degree be complemented by other sources of GGPP cannot be answered at the moment, because we do not have mutant(s) with null GGPPS activity in both the cytosol and mitochondria. What we can say so far is that the ggpps11 mutant has a less severe phenotype than the Rab geranylgeranylation null mutant [82]. A complete lack of Rab prenylation, which uses GGPP as a substrate, causes male sterility.
So far, in Arabidopsis, mutants in three genes synthesizing polyprenols in the cytosol and/or mitochondria were characterized, cpt1, cpt6, and lew1 (Table 2). In all three mutants, dolichol levels were reduced, but in each of them, different classes of dolichols were affected. While synthesis of long chain dolichols was reduced in the cpt1 and lew1 mutants [49], short chain dolichols were reduced in the cpt6 mutant [50]. The phenotype was also very different in all the mutants. No visible phenotype was reported for cpt6 [50], cpt1 null mutation affected plant growth and root elongation [49], and lew1 null mutation is lethal, although it is not known whether it is due to the embryo lethality or gametophyte lethality [88]. Mutants in later steps of dolichol biosynthesis, in polyprenol reductase (PPRD2), which catalyzes the conversion of polyprenol to dolichol, affects male gametophyte development and pollen germination [83]. Thus, the function of different CPTs in plant dolichol synthesis within the context of plant physiological processes still remains largely unknown.

4.2. Prenyltransferase Mutants in the Plastidial Branch of the Isoprenoid Pathway

The first product of DMAPP and IPP condensation in plastids is GPP, which in Arabidopsis is synthesized by the heteromeric enzyme GGPPS11/SSU (Figure 3). In agreement with that, both the knockdown ggpps11 mutant and the null ssu mutant have reduced levels of monoterpenoids to approximately 50% [11] (Table 2). Heterodimer GGPPS11/SSU is so far considered to be the only GPPS in Arabidopsis. Nevertheless, because monoterpenoids are still synthesized in Arabidopsis in the ssu knockout mutant, it is likely that other GPP-synthesizing enzymes are active in Arabidopsis. In addition to GGPPS11, small GPPS activity was also observed in GFPPSs [5,23] (Table S1). Previously, PPPS1 was classified as GPPS, because it was shown to synthesize GPP predominantly and was assumed to function in monoterpenoid biosynthesis and gibberellin biosynthesis in plastids [18,87]. Nevertheless, PPPS1-RNAi plants do not have a decreased level of monoterpenoids [8], and GPPS was shown to have PPPS (C20–C45) activity [41] (Table S1).
In agreement with the proposed housekeeping role of GGPPS11 [17,54], the ggpps11 knockdown mutant has lower levels of chlorophylls, carotenoids, phylloquinones, PQs, and tocopherols [8,36,54]. In addition to GGPPS11, which is the main GGPP-synthesizing enzyme in plastids and essential for the synthesis of photosynthesis-related isoprenoid compounds [54], GGPPS2 is another GGPPS paralog in plastids (Figure 3). The ggpps2 null mutant shows no visible developmental defects, and the level of carotenoids, chlorophylls, and prenylquinones is at the wild-type level (Table 2), demonstrating that GGPPS2 does not contribute significantly to the synthesis of photosynthesis-related plastidial isoprenoids that are essential for growth and development, and therefore, the function of this isozyme is yet to be determined [54].
The newly discovered group of PTs, GFPP synthases (GFPPS1-4) that synthesize GFPP, resides in plastids [17] (Figure 3). GFPPSs predominantly synthesize GFPP, which is a precursor of sesterterpenoids [5,23]. There are four GFPPSs in Arabidopsis plastids. GFPPS null mutants gfpps1, 2, 4 and GFPPS3-RNAi lines with 21% and 16% of mRNA levels, respectively, show no visible phenotypes [54] (Table 2).
Arabidopsis possesses two paralogous SPP synthases, SPPS1 and SPPS2, which assemble the side chain of PQ-9 in plastids (Figure 3). The spps1spps2 double knockout is devoid of PQ-9 and plastochromanol and cannot grow photoautotrophically [42] (Table 2). Because the preferred substrate for SPPS is GGPP and not DMAPP or GPP [59] (Table S1) and the ggpps11 mutant has decreased PQ-9 and plastochromanol levels [54], we can predict that the main enzyme generating substrate for SPPS in plastids is GGPPS11.
PPPS responsible for the synthesis of prenyl chain of UQ-9 in the mitochondria is also localized in plastids (PPPS1pl) because of double targeting [18,43] (Figure 3). While PPPS1mt synthesizes the side chain of UQ, PPPS1pl is not involved in PQ biosynthesis [43], and the function of PPPS1pl thus remains to be discovered. In tomato, downregulation of PPPS1 transcript led to a decreased level of gibberellins [87], but it was not tested whether PPPS1pl has the same function in Arabidopsis.
There is another PPPS in plastids, PPPS2, which synthesizes carbon chains of C20–40 [23] (Figure 3, Table S1). The null ppps2 mutant has no visible phenotype, and from all photosynthetically related metabolites that were analyzed, the level of chlorophyll b was slightly decreased and levels of β-carotene and neoxanthin slightly increased. The levels of other metabolites, such as chlorophyll a and phylloquinone were unchanged [54] (Table 2).
There is one CPT predicted to localize to plastids (Figure 3). The null mutant in plastidial CPT7 has a decreased level of plastidial polyprenols (C45–C55 with a maximum of C50) by 100% (Table 2). Affected plants have impaired photosystem II operating efficiency, and their thylakoids exhibit a decreased rate of electron transport but are not lethal [51].

5. Co-Expression Analysis and Prediction of Isoprenoid Fluxes

Co-expression analysis determines clusters of genes that are co-regulated together at expression levels. Metabolic pathway genes involved in the same physiological process are often co-regulated, which was also shown for genes in the isoprenoid pathway [89,90,91]. Therefore, gene association analysis based on the gene co-expression can be very instrumental in uncovering novel metabolic pathway genes or in determining their topology in the pathway gene network.
There are several examples where gene co-expression analysis resulted in better understanding of the prenyl-PP synthase function. For example, co-expression analysis with plastidial GGPPS and GFPPS isozymes showed that only GGPPS11, which is the ubiquitously expressed plastidial GGPPS paralog and the only one highly expressed in green tissues [17], is significantly co-expressed with all genes encoding MEP pathway enzymes and almost all genes from pathways synthesizing photosynthesis-related isoprenoids. The result of this co-expression analysis was also supported by the metabolite profiling of ggpps11 mutant [54] (Table 2). Co-expression analysis also correctly predicted subcellular localization of SPPS1 in plastids [42], although previous experimental evidence showed that this protein is associated with the ER [45,46]. This prompted further experiments that proved that SPPS1 is a paralog of SPPS2 and both enzymes are responsible for the plastidial synthesis of SPP and are essential for PQ and plastochromanol biosynthesis [42]. Co-expression analysis was also instrumental in assigning function to the PPPS1 protein. The PPPS1 protein was originally annotated as GPPS. It was shown that GPPS produces GPP in vitro, and GFP fusion of the protein (GPPS-GFP) localizes to plastids [18]. Nevertheless, the embryo lethal phenotype of the knockout mutant was rather nontypical for a plastidial isoprenoid pathway enzyme. In search of Arabidopsis isoprenoid pathway genes that are co-expressed with the Arabidopsis Ubi/Coq orthologs (UQ biosynthesis genes), GPPS was surprisingly found to be associated with the gene co-expression cluster. The enzymatic function of GPPS was reevaluated, and it was shown that GPPS actually synthesizes PPPs of different length (C25–C45), with C35 being the predominant prenyl-PP, and the enzyme was renamed to be PPP synthase 1 (PPPS1) [41,43]. Localization experiments using the expression of the GFP fusion protein driven by the native promoter have shown that PPPS1 localizes both to plastids and to the mitochondria [43].
We applied the same co-expression approach, which we used earlier for the plastidial GGPPS and GFPPS gene paralogs [54], to short trans-PTs to see whether we can extract additional information as to the function of PTs within the isoprenoid pathway gene network (Table 3, Table S2). Microarray expression data from different A. thaliana organs and tissues were used for analysis as described in Table S2. In the cytosol, as expected, FPPS1 and FPPS2 genes are the most co-expressed with the genes from the MVA and sterol biosynthetic pathways and both to the same extent supporting functional redundancy of both enzymes. As to the other functionally uncharacterized PTs and their involvement in the biosynthetic pathways in the cytosol and mitochondria, we can see that both FPPS and GGPPS genes and gene paralogs have connections to the genes of the UQ biosynthetic pathway and to the protein prenylation, and there is likely no strict allocation of a single FPPS or GGPPS gene paralog to those pathways. This seems not to be true for the diterpenoid biosynthesis, where GGPPS3 has the most gene connections to TPSs. In plastids, the synthesis of isoprenoid compounds associated with photosynthesis is associated with the single GGPPS11 gene paralog, as observed previously [54], except in the phylloquinone biosynthetic pathway, which is here, with the current gene expression set, more connected to GGPPS2 and other PTs. Genes synthesizing terpenoids and gibberellins are also instead associated with several isoprenoid synthases and their paralogs. Surprisingly, monoterpenoids show more connections to PPPS1, which was previously annotated as GPPS, than with SSU.
Metabolic gene clusters within dynamic chromosomal regions could provide another mechanism of transcriptional co-regulation. It has become clear in recent years that many plant secondary metabolic pathways or some of their genes are clustered [92]. These clusters have arisen by recruitment of genes from elsewhere in the genome through duplication and neofunctionalization. GFPPS–TPS gene clusters, where GFPPS genes and TPS genes are in close proximity were also found in Arabidopsis [23,56,57].

6. Conclusions

In the last decade, we have witnessed enormous progress in the characterization of Arabidopsis isoprenoid pathway PTs. Enzymes that were initially defined only by their approximate enzymatic function are now more precisely characterized at the enzymatic level, and their biological role, as well as topology, within the isoprenoid pathway network is clearly defined. PTs with essential function in the cytosol, mitochondria, and plastids are known as metabolic processes controlled by those PTs. What still remains as a challenge is to define at the gene level metabolic routes (fluxes) between PTs and TPSs and between PTs and pathways where prenyl-PP as a substrate is ambiguous; to understand minor, dynamic, developmentally controlled fluxes; and, last but not least, to explore more physiological functions of cis-PTs.

Supplementary Materials

The following are available online. Figure S1: Multiple sequence alignment for A. thaliana cis-PTs listed in Table 1; Table S1: Substrate and product specificity and kinetics of prenyltransferases in A. thaliana; Table S2: Gene expression data.

Author Contributions

Both authors contributed equally to this work.

Funding

This work was supported by grant VEGA 1/0926/17 from the Scientific grant agency of the Slovak Ministry of Education, Science, Research and Sport and of the Slovak Academy of Sciences.

Acknowledgments

We appreciate the help of Diana Coman Schmid (ETH Zurich) with co-expression analysis.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Wallrapp, F.H.; Pan, J.-J.; Ramamoorthy, G.; Almonacid, D.E.; Hillerich, B.S.; Seidel, R.; Patskovsky, Y.; Babbitt, P.C.; Almo, S.C.; Jacobson, M.P.; et al. Prediction of function for the polyprenyl transferase subgroup in the isoprenoid synthase superfamily. Proc. Natl. Acad. Sci. USA 2013, 110, E1196–E1202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Vranová, E.; Coman, D.; Gruissem, W. Network Analysis of the MVA and MEP Pathways for Isoprenoid Synthesis. Annu. Rev. Plant. Biol. 2013, 64, 665–700. [Google Scholar] [CrossRef] [PubMed]
  3. Liang, P.-H.; Ko, T.-P.; Wang, A.H.-J. Structure, mechanism and function of prenyltransferases. Eur. J. Biochem. 2002, 269, 3339–3354. [Google Scholar] [CrossRef] [PubMed]
  4. Winkelblech, J.; Shu-Ming Li, A.-F. Prenyltransferases as key enzymes in primary and secondary metabolism. Appl. Microbiol. Biotech. 2015, 99, 7379. [Google Scholar] [CrossRef]
  5. Nagel, R.; Bernholz, C.; Vranová, E.; Košuth, J.; Bergau, N.; Ludwig, S.; Wessjohann, L.; Gershenzon, J.; Tissier, A.; Schmidt, A. Arabidopsis thaliana isoprenyl diphosphate synthases produce the C25 intermediate geranylfarnesyl diphosphate. Plant. J. 2015, 84, 847–859. [Google Scholar] [CrossRef]
  6. Koyama, T.; Ogura, K. Isopentenyl Diphosphate Isomerase and Prenyltransferases. In Isoprenoids Including Carotenoids and Steroids; Cane, D.E., Ed.; Elsevier: Amsterdam, The Netherlands, 1999; Volume 2, pp. 69–95. [Google Scholar]
  7. Vandermoten, S.; Haubruge, É.; Cusson, M. New insights into short-chain prenyltransferases: Structural features, evolutionary history and potential for selective inhibition. Cell Mol. Life Sci. 2009, 66, 3685–3695. [Google Scholar] [CrossRef]
  8. Chen, Q.; Fan, D.; Wang, G. Heteromeric Geranyl(geranyl) Diphosphate Synthase Is Involved in Monoterpene Biosynthesis in Arabidopsis Flowers. Mol. Plant. 2015, 8, 1434–1437. [Google Scholar] [CrossRef] [Green Version]
  9. Wang, G.; Dixon, R.A. Heterodimeric geranyl(geranyl)diphosphate synthase from hop Humulus lupulus and the evolution of monoterpene biosynthesis. Proc. Natl. Acad. Sci. USA 2009, 106, 9914–9919. [Google Scholar] [CrossRef] [Green Version]
  10. Kwon, M.; Kwon, E.J.G.; Ro, D.K. Chapter Six - cis-Prenyltransferase and Polymer Analysis from a Natural Rubber Perspective. In Methods in Enzymology; O’Connor, S.E., Ed.; Elsevier: Amsterdam, The Netherlands, 2016; Volume 576, pp. 121–145. [Google Scholar] [CrossRef]
  11. Kumar, S.; Stecher, G.; Li, M.; Knyaz, C.; Tamura, K. MEGA X: Molecular Evolutionary Genetics Analysis across computing platforms. Mol. Biol. Evol. 2018, 35, 1547–1549. [Google Scholar] [CrossRef]
  12. Campbell, M.; Hahn, F.M.; Poulter, C.D.; Leustek, T. Analysis of the isopentenyl diphosphate isomerase gene family from Arabidopsis thaliana. Plant. Mol. Biol 1998, 36, 323–328. [Google Scholar] [CrossRef]
  13. Okada, K.; Kasahara, H.; Yamaguchi, S.; Kawaide, H.; Kamiya, Y.; Nojiri, H.; Yamane, H. Genetic evidence for the role of isopentenyl diphosphate isomerases in the mevalonate pathway and plant development in Arabidopsis. Plant. Cell Physiol. 2008, 49, 604–616. [Google Scholar] [CrossRef] [PubMed]
  14. Phillips, M.A.; D’Auria, J.C.; Gershenzon, J.; Pichersky, E. The Arabidopsis thaliana Type I Isopentenyl Diphosphate Isomerases Are Targeted to Multiple Subcellular Compartments and Have Overlapping Functions in Isoprenoid Biosynthesis. Plant. Cell 2008, 20, 677–696. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Sapir-Mir, M.; Mett, A.; Belausov, E.; Tal-Meshulam, S.; Frydman, A.; Gidoni, D.; Eyal, Y. Peroxisomal Localization of Arabidopsis Isopentenyl Diphosphate Isomerases Suggests That Part of the Plant Isoprenoid Mevalonic Acid Pathway Is Compartmentalized to Peroxisomes. Plant. Physiol. 2008, 148, 1219–1228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Vranová, E.; Hirsch-Hoffmann, M.; Gruissem, W. AtIPD: A Curated Database of Arabidopsis Isoprenoid Pathway Models and Genes for Isoprenoid Network Analysis. Plant. Physiol. 2011, 156, 1655–1660. [Google Scholar] [CrossRef] [Green Version]
  17. Beck, G.; Coman, D.; Herren, E.; Ruiz-Sola, M.Á.; Rodríguez-Concepción, M.; Gruissem, W.; Vranová, E. Characterization of the GGPP synthase gene family in Arabidopsis thaliana. Plant. Mol. Biol. 2013, 82, 393–416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Bouvier, F.; Suire, C.; d’Harlingue, A.; Backhaus, R.A.; Camara, B. Molecular cloning of geranyl diphosphate synthase and compartmentation of monoterpene synthesis in plant cells. Plant. J. 2000, 24, 241–252. [Google Scholar] [CrossRef]
  19. Joly, A.; Edwards, P.A. Effect of site-directed mutagenesis of conserved aspartate and arginine residues upon farnesyl diphosphate synthase activity. J. Biol. Chem. 1993, 268, 26983–26989. [Google Scholar]
  20. Song, L.; Poulter, C.D. Yeast farnesyl-diphosphate synthase: Site-directed mutagenesis of residues in highly conserved prenyltransferase domains I and II. Proc. Natl. Acad. Sci. USA 1994, 91, 3044–3048. [Google Scholar] [CrossRef] [Green Version]
  21. Armenteros, J.J.A.; Salvatore, M.; Winther, O.; Emanuelsson, O.; von Heijne, G.; Elofsson, A.; Nielsen, H. Detecting Sequence Signals in Targeting Peptides Using Deep Learning. Life Sci. Alliance 2019, 2, e201900429. [Google Scholar] [CrossRef] [Green Version]
  22. Hosfield, D.J.; Zhang, Y.; Dougan, D.R.; Broun, A.; Tari, L.W.; Swanson, R.V.; Finn, J. Structural Basis for Bisphosphonate-mediated Inhibition of Isoprenoid Biosynthesis. J. Biol. Chem. 2004, 279, 8526–8529. [Google Scholar] [CrossRef] [Green Version]
  23. Wang, C.; Chen, Q.; Fan, D.; Li, J.; Wang, G.; Zhang, P. Structural Analyses of Short-Chain Prenyltransferases Identify an Evolutionarily Conserved GFPPS Clade in Brassicaceae Plants. Mol. Plant. 2016, 9, 195–204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Lange, B.M.; Ghassemian, M. Genome organization in Arabidopsis thaliana: A survey for genes involved in isoprenoid and chlorophyll metabolism. Plant. Mol. Biol. 2003, 51, 925–948. [Google Scholar] [CrossRef] [PubMed]
  25. Keim, V.; Manzano, D.; Fernández, F.J.; Closa, M.; Andrade, P.; Caudepón, D.; Bortolotti, C.; Vega, M.C.; Arró, M.; Ferrer, A. Characterization of Arabidopsis FPS Isozymes and FPS Gene Expression Analysis Provide Insight into the Biosynthesis of Isoprenoid Precursors in Seeds. PLoS ONE 2012, 7, 49109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Running, M.P. The role of lipid post-translational modification in plant developmental processes. Front. Plant. Sci. 2014, 5, 50. [Google Scholar] [CrossRef] [Green Version]
  27. Delourme, D.; Lacroute, F.; Karst, F. Cloning of an Arabidopsis thaliana cDNA coding for farnesyl diphosphate synthase by functional complementation in yeast. Plant. Mol. Biol. 1994, 26, 1867–1873. [Google Scholar] [CrossRef]
  28. Cunillera, N.; Arró, M.; Delourme, D.; Karst, F.; Boronat, A.; Ferrer, A. Arabidopsis thaliana Contains Two Differentially Expressed Farnesyl-Diphosphate Synthase Genes. J. Biol. Chem. 1996, 271, 7774–7780. [Google Scholar] [CrossRef] [Green Version]
  29. Masferrer, A.; Arro, M.; Manzano, D.; Schaller, H.; Fernandez-Busquets, X.; Moncalean, P.; Fernandez, B.; Cunillera, N.; Boronat, A.; Ferrer, A. Overexpression of Arabidopsis thaliana farnesyl diphosphate synthase (FPS1S) in transgenic Arabidopsis induces a cell death/senescence-like response and reduced cytokinin levels. Plant. J. 2002, 30, 123–132. [Google Scholar] [CrossRef]
  30. Manzano, D.; Fernández-Busquets, X.; Schaller, H.; González, V.; Boronat, A.; Arró, M.; Ferrer, A. The metabolic imbalance underlying lesion formation in Arabidopsis thaliana overexpressing farnesyl diphosphate synthase (isoform 1S) leads to oxidative stress and is triggered by the developmental decline of endogenous HMGR activity. Planta 2004, 219, 982–992. [Google Scholar] [CrossRef]
  31. Closa, M.; Vranová, E.; Bortolotti, C.; Bigler, L.; Arró, M.; Ferrer, A.; Gruissem, W. The Arabidopsis thaliana FPP synthase isozymes have overlapping and specific functions in isoprenoid biosynthesis, and complete loss of FPP synthase activity causes early developmental arrest. Plant. J. 2010, 63, 512–525. [Google Scholar] [CrossRef]
  32. Manzano, D.; Andrade, P.; Caudepón, D.; Altabella, T.; Arró, M.; Ferrer, A. Suppressing Farnesyl Diphosphate Synthase Alters Chloroplast Development and Triggers Sterol-Dependent Induction of Jasmonate- and Fe-Related Responses. Plant. Physiol. 2016, 172, 93–117. [Google Scholar] [CrossRef]
  33. Bhatia, V.; Maisnam, J.; Jain, A.; Sharma, K.K.; Bhattacharya, R. Aphid-repellent pheromone E-β-farnesene is generated in transgenic Arabidopsis thaliana over-expressing farnesyl diphosphate synthase2. Ann. Bot 2015, 115, 581–591. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Okada, K.; Saito, T.; Nakagawa, T.; Kawamukai, M.; Kamiya, Y. Five geranylgeranyl diphosphate synthases expressed in different organs are localized into three subcellular compartments in Arabidopsis. Plant. Physiol. 2000, 122, 1045–1056. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Sakakibara, H.; Kasahara, H.; Ueda, N.; Kojima, M.; Takei, K.; Hishiyama, S.; Asami, T.; Okada, K.; Kamiya, Y.; Yamaya, T.; et al. Agrobacterium tumefaciens increases cytokinin production in plastids by modifying the biosynthetic pathway in the host plant. Proc. Natl. Acad. Sci. USA 2005, 102, 9972–9977. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Ruiz-Sola, M.Á.; Barja, M.V.; Manzano, D.; Llorente, B.; Schipper, B.; Beekwilder, J.; Rodriguez-Concepcion, M. A Single Arabidopsis Gene Encodes Two Differentially Targeted Geranylgeranyl Diphosphate Synthase Isoforms. Plant. Phys. 2016, 172, 1393–1402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Cunillera, N.; Boronat, A.; Ferrer, A. The Arabidopsis thaliana FPS1 Gene Generates a Novel mRNA That Encodes a Mitochondrial Farnesyl-diphosphate Synthase Isoform. J. Biol. Chem. 1997, 272, 15381–15388. [Google Scholar] [CrossRef] [Green Version]
  38. Zhu, X.; Suzuki, K.; Saito, T.; Okada, K.; Tanaka, K.; Nakagawa, T.; Matsuda, H.; Kawamukai, M. Geranylgeranyl pyrophosphate synthase encoded by the newly isolated gene GGPS6 from Arabidopsis thaliana is localized in mitochondria. Plant. Mol. Biol. 1997, 35, 331–341. [Google Scholar] [CrossRef]
  39. Lee, J.; Lee, H.; Kim, J.; Lee, S.; Kim, D.H.; Kim, S.; Hwang, I. Both the Hydrophobicity and a Positively Charged Region Flanking the C-Terminal Region of the Transmembrane Domain of Signal-Anchored Proteins Play Critical Roles in Determining Their Targeting Specificity to the Endoplasmic Reticulum or Endosymbiotic Organelles in Arabidopsis Cells. Plant. Cell 2011, 23, 1588. [Google Scholar] [CrossRef] [Green Version]
  40. Vranová, E.; Kopcsayová, D.; Košuth, J.; Colinas, M. Mutant-Based Model of Two Independent Pathways for Carotenoid-Mediated Chloroplast Biogenesis in Arabidopsis Embryos. Front. Plant. Sci. 2019, 10, 1034. [Google Scholar] [CrossRef]
  41. Hsieh, F.-L.; Chang, T.-H.; Ko, T.-P.; Wang, A.H.J. Structure and Mechanism of an Arabidopsis Medium/Long-Chain-Length Prenyl Pyrophosphate Synthase. Plant. Phys. 2011, 155, 1079–1090. [Google Scholar] [CrossRef] [Green Version]
  42. Block, A.; Fristedt, R.; Rogers, S.; Kumar, J.; Barnes, B.; Barnes, J.; Elowsky, C.G.; Wamboldt, Y.; Mackenzie, S.A.; Redding, K.; et al. Functional Modeling Identifies Paralogous Solanesyl-diphosphate Synthases That Assemble the Side Chain of Plastoquinone-9 in Plastids. J. Biol. Chem. 2013, 288, 27594–27606. [Google Scholar] [CrossRef] [Green Version]
  43. Ducluzeau, A.-L.; Wamboldt, Y.; Elowsky, C.G.; Mackenzie, S.A.; Schuurink, R.C.; Basset, G.J.C. Gene network reconstruction identifies the authentic trans-prenyl diphosphate synthase that makes the solanesyl moiety of ubiquinone-9 in Arabidopsis. Plant. J. 2012, 69, 366–375. [Google Scholar] [CrossRef] [PubMed]
  44. Kim, E.-H.; Lee, Y.; Kim, H.U. Fibrillin 5 Is Essential for Plastoquinone-9 Biosynthesis by Binding to Solanesyl Diphosphate Synthases in Arabidopsis. Plant. Cell 2015, 27, 2956–2971. [Google Scholar] [CrossRef] [PubMed]
  45. Jun, L.; Saiki, R.; Tatsumi, K.; Nakagawa, T.; Kawamukai, M. Identification and Subcellular Localization of Two Solanesyl Diphosphate Synthases from Arabidopsis thaliana. Plant. Cell Physiol. 2004, 45, 1882–1888. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Hirooka, K.; Izumi, Y.; An, C.-I.; Nakazawa, Y.; Fukusaki, E.; Kobayashi, A. Functional analysis of two solanesyl diphosphate synthases from Arabidopsis thaliana. Biosci. Biotechnol. Biochem. 2005, 69, 592–601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Surmacz, L.; Swiezewska, E. Polyisoprenoids—Secondary metabolites or physiologically important superlipids? Biochem. Bioph. Res. Co 2011, 407, 627–632. [Google Scholar] [CrossRef] [PubMed]
  48. Kera, K.; Takahashi, S.; Sutoh, T.; Koyama, T.; Nakayama, T. Identification and characterization of a cis,trans-mixed heptaprenyl diphosphate synthase from Arabidopsis thaliana. FEBS J. 2012, 279, 3813–3827. [Google Scholar] [CrossRef]
  49. Surowiecki, P.; Onysk, A.; Manko, K.; Swiezewska, E.; Surmacz, L. Long-Chain Polyisoprenoids Are Synthesized by AtCPT1 in Arabidopsis thaliana. Molecules 2019, 24, 2789. [Google Scholar] [CrossRef] [Green Version]
  50. Surmacz, L.; Plochocka, D.; Kania, M.; Danikiewicz, W.; Swiezewska, E. cis-Prenyltransferase AtCPT6 produces a family of very short-chain polyisoprenoids in planta. Biochim. Biophys. Acta 2014, 1841, 240–250. [Google Scholar] [CrossRef]
  51. Akhtar, T.A.; Surowiecki, P.; Siekierska, H.; Kania, M.; Van Gelder, K.; Rea, K.A.; Virta, L.K.A.; Vatta, M.; Gawarecka, K.; Wojcik, J.; et al. Polyprenols Are Synthesized by a Plastidial cis-Prenyltransferase and Influence Photosynthetic Performance. Plant. Cell 2017, 29, 1709–1725. [Google Scholar] [CrossRef] [Green Version]
  52. Hooper, C.M.; Castleden, I.R.; Tanz, S.K.; Aryamanesh, N.; Millar, A.H. SUBA4: The interactive data analysis centre for Arabidopsis subcellular protein locations. Nucleic Acids Res. 2016, 45, 1064–1074. [Google Scholar] [CrossRef] [Green Version]
  53. Zhu, X.; Suzuki, K.; Okada, K.; Tanaka, K.; Nakagawa, T.; Kawamukai, M.; Matsuda, K. Cloning and functional expression of a novel geranylgeranyl pyrophosphate synthase gene from Arabidopsis thaliana in Escherichia coli. Plant. Cell Physiol. 1997, 38, 357–361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Ruiz-Sola, M.Á.; Coman, D.; Beck, G.; Barja, M.V.; Colinas, M.; Graf, A.; Welsch, R.; Rütimann, P.; Bühlmann, P.; Bigler, L.; et al. Arabidopsis Geranylgeranyl Diphosphate Synthase 11 is a hub isozyme required for the production of most photosynthesis-related isoprenoids. New Phytol. 2016, 209, 252–264. [Google Scholar] [CrossRef] [PubMed]
  55. Evidente, A.; Kornienko, A.; Lefranc, F.; Cimmino, A.; Dasari, R.; Evidente, M.; Mathieu, V.; Kiss, R. Sesterterpenoids with Anticancer Activity. Curr. Med. Chem. 2015, 22, 3502–3522. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Huang, A.C.; Kautsar, S.A.; Hong, Y.J.; Medema, M.H.; Bond, A.D.; Tantillo, D.J.; Osbourn, A. Unearthing a sesterterpene biosynthetic repertoire in the Brassicaceae through genome mining reveals convergent evolution. Proc. Natl. Acad. Sci. USA 2017, 114, 6005–6014. [Google Scholar] [CrossRef] [Green Version]
  57. Shao, J.; Chen, Q.-W.; Lv, H.-J.; He, J.; Liu, Z.-F.; Lu, Y.-N.; Liu, H.-L.; Wang, G.-D.; Wang, Y. (+)-Thalianatriene and (−)-Retigeranin B Catalyzed by Sesterterpene Synthases from Arabidopsis thaliana. Org. Lett. 2017, 19, 1816–1819. [Google Scholar] [CrossRef]
  58. Nowicka, B.; Kruk, J. Occurrence, biosynthesis and function of isoprenoid quinones. Biochim. Biophys. Acta 2010, 1797, 1587–1605. [Google Scholar] [CrossRef] [Green Version]
  59. Hirooka, K.; Bamba, T.; Fukusaki, E.; Kobayashi, A. Cloning and kinetic characterization of Arabidopsis thaliana solanesyl diphosphate synthase. Biochem. J. 2003, 370, 679–686. [Google Scholar] [CrossRef]
  60. Mackenzie, S.A. Plant organellar protein targeting: A traffic plan still under construction. Trends Cell. Biol. 2005, 15, 548–554. [Google Scholar] [CrossRef] [Green Version]
  61. Liang, P.-H. Reaction Kinetics, Catalytic Mechanisms, Conformational Changes, and Inhibitor Design for Prenyltransferases. Biochemistry 2009, 48, 6562–6570. [Google Scholar] [CrossRef]
  62. Swiezewska, E.; Danikiewicz, W. Polyisoprenoids: Structure, biosynthesis and function. Prog. Lipid. Res. 2005, 44, 235–258. [Google Scholar] [CrossRef]
  63. Oh, S.K.; Han, K.H.; Ryu, S.B.; Kang, H. Molecular Cloning, Expression, and Functional Analysis of a cis-Prenyltransferase from Arabidopsis thaliana. Implications in Rubber Biosynthesis. J. Biol. Chem. 2000, 275, 18482–18488. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Cunillera, N.; Arro, M.; Fores, O.; Manzano, D.; Ferrer, A. Characterization of dehydrodolichyl diphosphate synthase of Arabidopsis thaliana, a key enzyme in dolichol biosynthesis. FEBS Lett. 2000, 477, 170–174. [Google Scholar] [CrossRef] [Green Version]
  65. Grabińska, K.A.; Edani, B.H.; Park, E.J.; Kraehling, J.R.; Sessa, W.C. A conserved C-terminal RXG motif in the NgBR subunit of cis-prenyltransferase is critical for prenyltransferase activity. J. Biol. Chem. 2017, 292, 17351–17361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Qu, Y.; Chakrabarty, R.; Tran, H.T.; Kwon, E.-J.G.; Kwon, M.; Nguyen, T.-D.; Ro, D.-K. A lettuce (Lactuca sativa) homolog of human Nogo-B receptor interacts with cis-prenyltransferase and is necessary for natural rubber biosynthesis. J. Biol. Chem. 2015, 290, 1898–1914. [Google Scholar] [CrossRef] [Green Version]
  67. Epping, J.; van Deenen, N.; Niephaus, E.; Stolze, A.; Fricke, J.; Huber, C.; Eisenreich, W.; Twyman, R.M.; Prüfer, D.; Schulze Gronover, C. A rubber transferase activator is necessary for natural rubber biosynthesis in dandelion. Nat. Plants 2015, 1, 15048. [Google Scholar] [CrossRef]
  68. Brasher, M.I.; Surmacz, L.; Leong, B.; Pitcher, J.; Swiezewska, E.; Pichersky, E.; Akhtar, T.A. A two-component enzyme complex is required for dolichol biosynthesis in tomato. Plant. J. 2015, 82, 903–914. [Google Scholar] [CrossRef] [Green Version]
  69. Addou, S.; Rentzsch, R.; Lee, D.; Orengo, C.A. Domain-Based and Family-Specific Sequence Identity Thresholds Increase the Levels of Reliable Protein Function Transfer. J. Mol. Biol. 2009, 387, 416–430. [Google Scholar] [CrossRef]
  70. Chen, A.; Kroon, P.A.; Poulter, C.D. Isoprenyl diphosphate synthases: Protein sequence comparisons, a phylogenetic tree, and predictions of secondary structure. Protein Sci. 1994, 3, 600–607. [Google Scholar] [CrossRef] [Green Version]
  71. Tarshis, L.C.; Proteau, P.J.; Kellogg, B.A.; Sacchettini, J.C.; Poulter, C.D. Regulation of product chain length by isoprenyl diphosphate synthases. Proc. Natl. Acad. Sci. USA 1996, 93, 15018–15023. [Google Scholar] [CrossRef] [Green Version]
  72. Ohnuma, S.-i.; Hirooka, K.; Tsuruoka, N.; Yano, M.; Ohto, C.; Nakane, H.; Nishino, T. A Pathway Where Polyprenyl Diphosphate Elongates in Prenyltransferase: insight into a common mechanism of chain length determination of prenyltransferases. J. Biol. Chem. 1998, 273, 26705–26713. [Google Scholar] [CrossRef] [Green Version]
  73. Guo, R.-T.; Kuo, C.-J.; Chou, C.-C.; Ko, T.-P.; Shr, H.-L.; Liang, P.-H.; Wang, A.H.J. Crystal Structure of Octaprenyl Pyrophosphate Synthase from Hyperthermophilic Thermotoga maritima and Mechanism of Product Chain Length Determination. J. Biol. Chem. 2004, 279, 4903–4912. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Chang, T.-H.; Guo, R.-T.; Ko, T.-P.; Wang, A.H.J.; Liang, P.-H. Crystal Structure of Type-III Geranylgeranyl Pyrophosphate Synthase from Saccharomyces cerevisiae and the Mechanism of Product Chain Length Determination. J. Biol. Chem. 2006, 281, 14991–15000. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Gabelli, S.B.; McLellan, J.S.; Montalvetti, A.; Oldfield, E.; Docampo, R.; Amzel, L.M. Structure and mechanism of the farnesyl diphosphate synthase from Trypanosoma cruzi: Implications for drug design. Proteins 2006, 62, 80–88. [Google Scholar] [CrossRef] [PubMed]
  76. Kavanagh, K.L.; Dunford, J.E.; Bunkoczi, G.; Russell, R.G.G.; Oppermann, U. The Crystal Structure of Human Geranylgeranyl Pyrophosphate Synthase Reveals a Novel Hexameric Arrangement and Inhibitory Product Binding. J. Biol. Chem. 2006, 281, 22004–22012. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Kulmanov, M.; Khan, M.A.; Hoehndorf, R.; Wren, J. DeepGO: Predicting protein functions from sequence and interactions using a deep ontology-aware classifier. Bioinformatics 2018, 34, 660–668. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Bepler, T.; Berger, B. Learning Protein Sequence Embeddings Using Information from Structure. In Proceedings of the International Conference on Learning Representations, New Orleans, LA, USA, 6–9 May 2019. [Google Scholar]
  79. Jin, H.; Song, Z.; Nikolau, B.J. Reverse genetic characterization of two paralogous acetoacetyl CoA thiolase genes in Arabidopsis reveals their importance in plant growth and development. Plant. J. 2012, 70, 1015–1032. [Google Scholar] [CrossRef]
  80. Okada, K.; Ohara, K.; Yazaki, K.; Nozaki, K.; Uchida, N.; Kawamukai, M.; Nojiri, H.; Yamane, H. The AtPPT1 gene encoding 4-hydroxybenzoate polyprenyl diphosphate transferase in ubiquinone biosynthesis is required for embryo development in Arabidopsis thaliana. Plant. Mol. Biol 2004, 55, 567–577. [Google Scholar] [CrossRef]
  81. Babiychuk, E.; Bouvier-Navé, P.; Compagnon, V.; Suzuki, M.; Muranaka, T.; Van Montagu, M.; Kushnir, S.; Schaller, H. Allelic mutant series reveal distinct functions for Arabidopsis cycloartenol synthase 1 in cell viability and plastid biogenesis. Proc. Natl. Acad. Sci. USA 2008, 105, 3163–3168. [Google Scholar] [CrossRef] [Green Version]
  82. Gutkowska, M.; Wnuk, M.; Nowakowska, J.; Lichocka, M.; Stronkowski, M.M.; Swiezewska, E. Rab geranylgeranyl transferase β subunit is essential for male fertility and tip growth in Arabidopsis. J. Exp. Bot 2015, 66, 213–224. [Google Scholar] [CrossRef]
  83. Jozwiak, A.; Gutkowska, M.; Gawarecka, K.; Surmacz, L.; Buczkowska, A.; Lichocka, M.; Nowakowska, J.; Swiezewska, E. POLYPRENOL REDUCTASE2 Deficiency Is Lethal in Arabidopsis Due to Male Sterility. Plant. Cell 2015, 27, 3336–3353. [Google Scholar] [CrossRef]
  84. Motohashi, R.; Ito, T.; Kobayashi, M.; Taji, T.; Nagata, N.; Asami, T.; Yoshida, S.; Yamaguchi-Shinozaki, K.; Shinozaki, K. Functional analysis of the 37 kDa inner envelope membrane polypeptide in chloroplast biogenesis using a Ds-tagged Arabidopsis pale-green mutant. Plant. J. 2003, 34, 719–731. [Google Scholar] [CrossRef] [PubMed]
  85. Yamasato, A.; Tanaka, R.; Tanaka, A. Loss of the N-terminal domain of chlorophyllide a oxygenase induces photodamage during greening of Arabidopsis seedlings. BMC Plant. Biol. 2008, 8, 64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Ruppel, N.J.; Kropp, K.N.; Davis, P.A.; Martin, A.E.; Luesse, D.R. Mutations in Geranylgeranyl Diphosphate Synthase 1 affect chloroplast development in Arabidopsis thaliana (Brassicaceae) Am. J. Bot 2013, 100, 2074–2084. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. van Schie, C.C.N.; Ament, K.; Schmidt, A.; Lange, T.; Haring, M.A.; Schuurink, R.C. Geranyl diphosphate synthase is required for biosynthesis of gibberellins. Plant. J. 2007, 52, 752–762. [Google Scholar] [CrossRef]
  88. Zhang, H.; Ohyama, K.; Boudet, J.; Chen, Z.; Yang, J.; Zhang, M.; Muranaka, T.; Maurel, C.; Zhu, J.-K.; Gong, Z. Dolichol Biosynthesis and Its Effects on the Unfolded Protein Response and Abiotic Stress Resistance in Arabidopsis. Plant. Cell 2008, 20, 1879–1898. [Google Scholar] [CrossRef] [Green Version]
  89. Wille, A.; Zimmermann, P.; Vranova, E.; Furholz, A.; Laule, O.; Bleuler, S.; Hennig, L.; Prelic, A.; von Rohr, P.; Thiele, L.; et al. Sparse graphical Gaussian modeling of the isoprenoid gene network in Arabidopsis thaliana. Genome Biol. 2004, 5, 92. [Google Scholar] [CrossRef] [Green Version]
  90. Ghassemian, M.; Lutes, J.; Tepperman, J.M.; Chang, H.-S.; Zhu, T.; Wang, X.; Quail, P.H.; Markus Lange, B. Integrative analysis of transcript and metabolite profiling data sets to evaluate the regulation of biochemical pathways during photomorphogenesis. Arch. Biochem. Biophys. 2006, 448, 45–59. [Google Scholar] [CrossRef]
  91. Meier, S.; Tzfadia, O.; Vallabhaneni, R.; Gehring, C.; Wurtzel, E. A transcriptional analysis of carotenoid, chlorophyll and plastidial isoprenoid biosynthesis genes during development and osmotic stress responses in Arabidopsis thaliana. BMC Syst. Biol. 2011, 5, 77. [Google Scholar] [CrossRef] [Green Version]
  92. Nützmann, H.-W.; Scazzocchio, C.; Osbourn, A. Metabolic Gene Clusters in Eukaryotes. Annu. Rev. Genet. 2018, 52, 159–183. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the reactions catalyzed by trans-prenyltransferases and cis-prenyltransferases showing the stereochemistry of the double bond formed in the enzyme product. DMAPP, dimethylallyl diphosphate; IPP, isopentenyl diphosphate; PT, prenyltransferase; GPP, geranyl diphosphate; FPP, farnesyl diphosphate; GGPP, geranylgeranyl diphosphate; GFPP, geranylfarnesyl diphosphate; and OPP, diphosphate ester.
Figure 1. Schematic representation of the reactions catalyzed by trans-prenyltransferases and cis-prenyltransferases showing the stereochemistry of the double bond formed in the enzyme product. DMAPP, dimethylallyl diphosphate; IPP, isopentenyl diphosphate; PT, prenyltransferase; GPP, geranyl diphosphate; FPP, farnesyl diphosphate; GGPP, geranylgeranyl diphosphate; GFPP, geranylfarnesyl diphosphate; and OPP, diphosphate ester.
Molecules 24 04556 g001
Figure 2. Phylogenetic analysis of the prenyltransferase family in A. thaliana. The phylogenetic tree illustrates the relatedness of PTs, which cluster into two major clades—trans- and cis-PTs. Enzyme abbreviations are listed in Table 1. Enzymes in white boxes are not PTs, but they interact with PTs—small subunit (SSU) with geranylgeranyl diphosphate synthase 11 (GGPPS11) [8,9] and LEW1 with cis-PT3 (CPT3) [10]. Because the tree was left unrooted, the branching order may not always reflect the evolutionary course that the enzymes have taken. The evolutionary analysis was conducted in MEGA X [11] using the neighbor-joining method with 500 bootstrap replicates. The scale bar represents 0.2 amino acid substitutions per site.
Figure 2. Phylogenetic analysis of the prenyltransferase family in A. thaliana. The phylogenetic tree illustrates the relatedness of PTs, which cluster into two major clades—trans- and cis-PTs. Enzyme abbreviations are listed in Table 1. Enzymes in white boxes are not PTs, but they interact with PTs—small subunit (SSU) with geranylgeranyl diphosphate synthase 11 (GGPPS11) [8,9] and LEW1 with cis-PT3 (CPT3) [10]. Because the tree was left unrooted, the branching order may not always reflect the evolutionary course that the enzymes have taken. The evolutionary analysis was conducted in MEGA X [11] using the neighbor-joining method with 500 bootstrap replicates. The scale bar represents 0.2 amino acid substitutions per site.
Molecules 24 04556 g002
Figure 3. Topology of the prenyltransferase genes within the isoprenoid pathway of A. thaliana. Based on the isoprenoid pathway network constructed by Vranová et al. [16]. Abbreviations of the enzymes in boxes are given in Table 1. MVA, mevalonate; MEP, 2-C-methyl-D-erythritol 4-phosphate; IPPI, isopentenyl diphosphate (IPP, C5) isomerase; IPPI1S, 2S, short IPPI variants; IPPI1L, 2L, long IPPI variants; DMAPP (C5), dimethylallyl diphosphate; GPP (C10), geranyl diphosphate; FPP (C15), farnesyl diphosphate; GGPP (C20), geranylgeranyl diphosphate; GFPP (C25), geranylfarnesyl diphosphate; PPP (C ≥ 30), polyprenyl diphosphate; SPP (C45), solanesyl diphosphate; PQ-9, plastoquinone 9; and UQ-9, ubiquinone 9.
Figure 3. Topology of the prenyltransferase genes within the isoprenoid pathway of A. thaliana. Based on the isoprenoid pathway network constructed by Vranová et al. [16]. Abbreviations of the enzymes in boxes are given in Table 1. MVA, mevalonate; MEP, 2-C-methyl-D-erythritol 4-phosphate; IPPI, isopentenyl diphosphate (IPP, C5) isomerase; IPPI1S, 2S, short IPPI variants; IPPI1L, 2L, long IPPI variants; DMAPP (C5), dimethylallyl diphosphate; GPP (C10), geranyl diphosphate; FPP (C15), farnesyl diphosphate; GGPP (C20), geranylgeranyl diphosphate; GFPP (C25), geranylfarnesyl diphosphate; PPP (C ≥ 30), polyprenyl diphosphate; SPP (C45), solanesyl diphosphate; PQ-9, plastoquinone 9; and UQ-9, ubiquinone 9.
Molecules 24 04556 g003
Figure 4. Multiple sequence alignment for A. thaliana trans-PTs, listed in Table 1. N-terminal sequences are omitted according to prediction by TargetP-2.0 [21] (http://www.cbs.dtu.dk/services/TargetP). Highly conserved regions are highlighted in gray. Only regions II and III are present in small subunit (SSU). Motifs essential for catalysis and binding of allylic substrate (first aspartate-rich motif (FARM), second aspartate-rich motif (SARM), and +5 and +6 arginines (RR)) are highlighted in black [19,20]. Basal residues active in IPP binding are marked with arrows [22]. The CXXXC motif highlighted in blue is critical for physical interaction between the large subunit (LSU) and the SSU [9]. The floor 1, 2, and 3 residues of the three-floor model [23] are indicated in green, red, and yellow, respectively. Blocking residues are boxed.
Figure 4. Multiple sequence alignment for A. thaliana trans-PTs, listed in Table 1. N-terminal sequences are omitted according to prediction by TargetP-2.0 [21] (http://www.cbs.dtu.dk/services/TargetP). Highly conserved regions are highlighted in gray. Only regions II and III are present in small subunit (SSU). Motifs essential for catalysis and binding of allylic substrate (first aspartate-rich motif (FARM), second aspartate-rich motif (SARM), and +5 and +6 arginines (RR)) are highlighted in black [19,20]. Basal residues active in IPP binding are marked with arrows [22]. The CXXXC motif highlighted in blue is critical for physical interaction between the large subunit (LSU) and the SSU [9]. The floor 1, 2, and 3 residues of the three-floor model [23] are indicated in green, red, and yellow, respectively. Blocking residues are boxed.
Molecules 24 04556 g004aMolecules 24 04556 g004b
Table 1. Enzymes and gene models of prenyltransferases in A. thaliana.
Table 1. Enzymes and gene models of prenyltransferases in A. thaliana.
ENZYMEGENE
AcronymNameEC NumberAGIGene ModelReferencesAlternative NamesReferencesLocalization bReferences
GPPSGeranyl diphosphate synthase cEC 2.5.1.1At4g36810/At4g38460GGPPS11La/SSUIDS11; GGPPS11; GGPS1/GGPPS12[5,24,34]/[24]Pl/Pl[17,34,35,36]/[8]
FPPSFarnesyl diphosphate synthaseEC 2.5.1.10At5g47770FPPS1Sa[24,37]C[25]
FPPS1La[24,37]Mt[37]
At4g17190FPPS2[24]C[25]
GGPPSGeranylgeranyl diphosphate synthaseEC 2.5.1.29At1g49530GGPPS1[24]IDS1; GGPS6[5,34]Mt[5,17,34,35,38]
At2g18620GGPPS2[24]IDS2[5]Pl[17]
At2g18640GGPPS3[24]IDS3; GGPS4[5,34]ER[17,34]
At2g23800GGPPS4[24]IDS4; GGPS2[5,34]ER[17,34,39]
At4g36810GGPPS11Sa[36]C[36,40]
At4g36810GGPPS11LaIDS11; GGPPS11; GGPS1[5,24,34]Pl[17,34,35]
GFPPSGeranylfarnesyl diphosphate synthaseEC 2.5.1.81At3g14530GFPPS1[23]IDS6; GGPPS6[5,24]Pl[5,17]
At3g14550GFPPS2[23]IDS7; GGPPS7; GGPS3[5,24,34]Pl[5,17,34]
At3g29430GFPPS3[23]IDS9; GGPPS9[5,24]Pl[5,17]
At3g32040GFPPS4[23]IDS10; GGPPS10[5,24]Pl[5,17]
PPPSPolyprenyl diphosphate synthaseEC 2.5.1.91At2g34630PPPS1[41]GPPS; SPPS3[24,42]Pl/Mt[18,43]
At3g20160PPPS2[23]IDS8; GGPPS8[5,24]Pl[17]
SPPSSolanesyl diphosphate synthaseEC 2.5.1.85At1g78510SPPS1[42]Pl[42,44]
At1g17050SPPS2[42]Pl[44,45,46]
CPTcis-PrenyltransferaseEC 2.5.1.87At1g11755LEW1[47]PMd
At2g23410CPT1[47]AtcPT3[48]ER[48,49]
At2g23400CPT2[47]AtcPT2[48]Cd
At2g17570CPT3[47]AtcPT1[48]Mtd
At5g60510CPT4[47]AtcPT9[48]Cd
At5g60500CPT5[47]AtcPT8[48]Cd
At5g58780CPT6[47]AtcPT5[48]ER[48,50]
At5g58770CPT7[47]AtcPT4/AtHEPS[48,51]Pl[48,51]
At5g58782CPT8[47]AtcPT6[48]ER[47]
At5g58784CPT9[47]AtcPT7[48]ERd
Abbreviations: EC, Enzyme Commission; AGI, The Arabidopsis Genome Initiative. a Alternative transcription start site. b Subcellular localization: C, cytosol; ER, endoplasmic reticulum; Mt, mitochondrion; Pl, plastid; PM, plasma membrane. c Heterodimeric GPP synthase. d SUBA prediction [52]; http://suba.live/aboutSUBA4.html.
Table 2. Mutations in prenyltransferase genes and their phenotypes in A. thaliana.
Table 2. Mutations in prenyltransferase genes and their phenotypes in A. thaliana.
GeneAGIMutagenAlleleMutant LinePhenotypeReferences
FPPS1At5g47770T-DNAfpps1-1SAIL_310-D07Wild-type phenotype with slightly reduced levels of sterols and UQ-9[31]
T-DNAfpps1-2SALK_073576
FPPS2At4g17190T-DNAfpps2-1SAIL_328_G06Wild-type phenotype with slightly reduced levels of sterols and UQ-9[31]
Dsfpps2-2GT7041
FPPS1/FPPS2At5g47770/At4g17190T-DNA/Dsfpps1/fpps2SAIL_310_D07; SALK_073576/SAIL_328_G06; GT7041Embryo lethal at the globular stage; slightly impaired pollen tube elongation[31]
amiRNAfpps1/fpps2amiFPS1 (21%/26% mRNA), amiFPS2 (16%/35% mRNA)Impaired growth and development; chlorosis; reduces level of chlorophylls, carotenoids, sterols and UQ-9; higher level of UQ-10[32]
SSUAt4g38460T-DNAssu-1 (ggpps12-1)pst11416Wild-type phenotype with reduced level of monoterpenoids, and wild-type level of carotenoids, chlorophylls, sesquiterpenoids[8]
T-DNAssu (ggpps12)pst01798 (20% mRNA)
GGPPS1At1g49530T-DNAggpps1-1SAIL_559_G01Wild-type phenotype with wild-type level of UQ-9, carotenoids, tocopherols, chlorophylls, PQ-9, phylloquinones, plastochromanol-8[36]
GGPPS2At2g18620T-DNAggpps2-1FLAG_134_B10Wild-type phenotype with wild-type levels of carotenoids, chlorophylls, phylloquinones[54]
GGPPS3At2g18640-----
GGPPS4At2g23800-----
GGPPS11At4g36810EMS mutagenesis ggpps11-1D163R point mutationVariegated phenotype; germination delayed on the inhibitor of gibberellin biosynthesis; reduced level of chlorophylls and carotenoids[54,86]
T-DNA in chloroplast targeting sequenceggpps11-2SALK_015098Albino seedling[54,86]
T-DNAggpps11-3SALK_085914Embryo lethal at the heart stage[54,86]
T-DNAggpps11-4SAIL_712_D06[54]
T-DNA in 5′-UTRggpps11-5SALK_140601Pale green phenotype with reduced level of monoterpenoids, carotenoids, tocopherols, chlorophylls, PQ-9, phylloquinones, plastochromanol-8[8,36,54]
GFPPS1At3g14530T-DNAgfpps1-1 (ggpps6-1)SAIL_1148_A03Wild-type phenotype with wild-type level of carotenoids, chlorophylls, phylloquinones[54]
GFPPS2At3g14550T-DNAgfpps2-1 (ggpps7-1)SALK_119280Wild-type phenotype with wild-type level of carotenoids, chlorophylls, phylloquinones[54]
GFPPS3At3g29430RNAigfpps3 (ggpps9)RNAi GGPPS9-1 (21% mRNA); RNAi GGPPS9-6 (16% mRNA)Wild-type phenotype with wild-type level of carotenoids, chlorophylls, phylloquinones[54]
GFPPS4At3g32040T-DNAgfpps4-1 (ggpps10-1)SM_3_32015Wild-type phenotypes with wild-type level of carotenoids, chlorophylls, phylloquinones[54]
PPPS1At2g34630T-DNAppps1-1 (gps1-1)GABI 097_G02Embryo lethal[87]
RNAippps1RNAi PPPS1-1-6 (10% mRNA)Growth reduction; dwarfed plants with delayed flowering[87]
RNAippps1RNAi PPPS1-1-3 (20% mRNA)Reduced level of UQ-9; wild-type level of PQ[43]
RNAippps1GPPS RNAi-2, GPPS RNAi-4 (10% mRNA)Wild-type phenotype with wild-type level of monoterpenoids and sesquiterpenoids[8]
PPPS2At3g20160T-DNAppps2-1 (ggpps8-1)FLAG_470_E09Higher level of carotenoids; reduced level of chlorophyll b[54]
SPPS1At1g78510T-DNAspps1SALK_126948Wild-type phenotype with reduced level of PQ and plastochromanol, and wild-type level of tocopherol and UQ[42]
SPPS2At1g17050T-DNAspps2SALK_064292Developmental delay; stunted phenotype and pale yellowish leaves at the high light; reduced level of PQ; no plastochromanol; higher level of tocopherol, and wild-type level of UQ[42]
SPPS1/SPPS2At1g78510/T-DNAspps1/spps2SALK_126948/SALK_064292Seedling lethal, albino phenotype; no PQ and plastochromanol; wild-type level of UQ and tocopherol[42]
At1g17050
LEW1At1g11755EMS mutagenesislew1-1G159A point mutationLeaf-wilting phenotype, increased drought resistance, impaired plasma membrane integrity, impaired protein N-glycosylation, reduced the total plant content of main dolichols C75, C80 by 85% and protein glycosylation defects[88]
T-DNAlew1-2SALK_032276Lethal[88]
CPT1At2g23410T-DNAcpt1-1SALK_038151Extremely stunted growth and shorter root length; reduced level from Dol-18 to -23 (Dol-21 dominating)[49]
T-DNAcpt1-2SALK_032276
T-DNAcpt1-3SALK_100795
CPT2At2g23400-----
CPT3At2g17570-----
CPT4At5g60510-----
CPT5At5g60500-----
CPT6At5g58780T-DNAcpt6-1SALK_071255Reduced level of Dol-7 and short-chain dolichols (Dol-13 dominating)[50]
cpt6-2SALK_064499
CPT7At5g58770T-DNAcpt7-1SALK_022111Decreased thylakoid membrane fluidity and photosynthetic performance; no polyprenols 9, 10, 11 (Pol-10 dominating); wild-type level of tocopherols, phylloquinone, carotenoids, PQ, and chlorophylls[51]
RNAicpt7RNAi-23; RNAi-24 and RNAi-31 (15% mRNA)
CPT8At5g58782-----
CPT9At5g58784-----
Table 3. trans-Prenyltransferase genes and isoprenoid pathway gene co-expressions in A. thaliana.
Table 3. trans-Prenyltransferase genes and isoprenoid pathway gene co-expressions in A. thaliana.
CYTOSOL AND MITOCHONDRIA
PathwayMVASterolTriterpenoidsSesquiterpenoids/DiterpenoidsProtein PrenylationUbiquinone
Total No. of Genes per Pathway924131374
FPPS1At5g477706132140
FPPS2At4g171907140131
GGPPS1At1g49530002003
GGPPS3At2g18640043612
GGPPS11At4g368100112010
GGPPS4At2g23800223102
PPPS1At2g34630230214
PLASTIDS
PathwayMEPPlastoquinoneChlorophyllCarotenoidPhylloquinoneGibberellinsMonoterpenoidsDiterpenoids/SesquiterpenoidsSesterterpenoids
Total No. of Genes per Pathway7637327236133
SSUAt4g3846055252205210
GGPPS11At4g3681073291817100
GGPPS2At2g18620003152271
GFPPS1At3g14530001032052
GGPPS3At3g29430004042281
GGPPS4At3g32040003057271
PPPS2At3g20160004044281
PPPS1At2g34630000242421
SPPS1At1g7851066292004210
SPPS2At1g1705074292207210
Molecules 24 04556 i001Table 3 shows number of pathway genes that are significantly co-expressed with the trans-PTs. The intensity of the highlighted background is directly proportional to the relative number of co-expressed genes/total number of pathway genes. Data for co-expression analysis were obtained from BAR/Expression Browser (http://bar.utoronto.ca/affydb/cgi-bin/affy_db_exprss_browser_in.cgi) and log2 transformed (see Table S2 for log2 transformed expression data). In general, AtGeneExpress_plus-Extended Tissue Series data/Average of replicate treatments were downloaded. Genes that are missing were not present on the microarrays, and therefore, they are not included in the analysis. Pearson correlation coefficients and the corresponding false discovery rate (FDR) p-values were calculated. The threshold for significance is p-value ≤ 0.05. A detailed description of the analysis can be found in [54]. Genes in categories mono-, di-, sesqui-, sester- and triterpenoids contain genes encoding only terpene cyclases. Genes in the category diterpenoids/sesquiterpenoids and sesquiterpenoids/diterpenoids partly overlap. In both categories were kept the genes that have no experimentally proven one or other activity, and plastidial genes were considered diterpenoids and kept only in the category diterpenoids/sesquiterpenoids.

Share and Cite

MDPI and ACS Style

Kopcsayová, D.; Vranová, E. Functional Gene Network of Prenyltransferases in Arabidopsis thaliana. Molecules 2019, 24, 4556. https://doi.org/10.3390/molecules24244556

AMA Style

Kopcsayová D, Vranová E. Functional Gene Network of Prenyltransferases in Arabidopsis thaliana. Molecules. 2019; 24(24):4556. https://doi.org/10.3390/molecules24244556

Chicago/Turabian Style

Kopcsayová, Diana, and Eva Vranová. 2019. "Functional Gene Network of Prenyltransferases in Arabidopsis thaliana" Molecules 24, no. 24: 4556. https://doi.org/10.3390/molecules24244556

Article Metrics

Back to TopTop