Functional and structural studies of the Presequence , PreP

Licentiate thesis

Hans G Bäckman

Supervisor: Elzbieta Glaser

Department of Biochemistry and Biophysics

Stockholm University

2014

2

Contents

Abstract ...... 5 Publications ...... 6 The thesis is based on the following publications ...... 6 Abbreviations ...... 7 Mitochondria and chloroplasts ...... 8 The mitochondrial genome ...... 9 Function and structure of mitochondria ...... 9 The chloroplast genome ...... 10 Function and structure of chloroplasts ...... 10 targeting and import into the mitochondrial matrix ...... 11 Mitochondrial processing peptidases ...... 13 Protein targeting and import into chloroplast stroma ...... 14 Stromal processing peptidase ...... 16 Dual-targeting to mitochondria and chloroplasts ...... 16 Proteolysis in mitochondria and chloroplasts ...... 17 The ATP dependent ...... 17 Clp proteases ...... 17 FtsH proteases ...... 17 Lon proteases ...... 18 The ATP independent proteases ...... 19 Rhomboid proteases ...... 19 Deg proteases ...... 19 OOP ...... 20 PreP ...... 20 Material & Methods ...... 21 Paper I...... 21 Generation of mutant Arabidopsis plants, single and double knockouts ...... 21 Growth conditions of A. thaliana plants ...... 21 Purification of mitochondria and chloroplasts ...... 21 Degradation assays ...... 22 Respiratory measurements ...... 22

3

Protein gel and immunoblot analysis of thylakoid membranes ...... 22 Protein extractions and analysis of AtPreP1 and AtPreP2 content ...... 22 Biomass...... 23 Paper II...... 23 Generation of AtPreP1 variants ...... 23 Expression and purification of AtPreP1 and its variants ...... 23 Degradation assays ...... 24 Paper III...... 24 Generation of hPreP-SNP variants ...... 24 Degradation assays ...... 24 Summary of results ...... 25 The Presequence protease, PreP ...... 25 Paper I - Deletion of an organellar peptidasome PreP affects early development in Arabidopsis thaliana ...... 26 Paper II - Binding of divalent cations is essential for the activity of the organellar peptidasome in Arabidopsis thaliana, AtPreP ...... 28 Paper III - Biochemical studies of SNPs of the mitochondrial Aβ-degrading protease, hPreP ...... 29 Future perspectives ...... 31 Structure and catalytic function of AtPreP1 ...... 31 Acknowledgements ...... 32 References ...... 33

4

Abstract

AtPreP (Arabidopsis thaliana Presequence Protease) is a zink metallooligopeptidase that is dually targeted to both mitochondria and chloroplasts. In these organelles it functions as a peptidasome that degrades the N-terminal targeting peptides that are cleaved off from the mature protein after protein import, as well as other unstructured peptides. In A. thaliana there are two isoforms of PreP, AtPreP1 and AtPreP2.

We have performed characterization studies of single and double prep knockout plants. Immunoblot analysis revealed that both PreP isoforms are expressed in all tissues with highest expression levels in flowers and siliques. Furthermore, AtPreP1 was shown to be the most abundant isoform of the two. When comparing phenotype, the atprep2 mutant was similar to wild type, whereas the atprep1 mutant had a slight pale-green phenotype in the early developmental stages. The atprep1 atprep2 double knockout plants showed a chlorotic phenotype in true leaves, especially prominent during the early developmental stages. When analysing the first true leaves of double knockout plants, we found a significant decrease in chlorophyll a and b content. Mitochondrial respiratory rates measurements showed partially uncoupled mitochondria. Ultrastructure analysis using electron microscopy on double knockout plants showed aberrant chloroplasts with altered grana stacking and clearly fewer starch granules. Older plants showed less altered phenotype, although there was a significant decrease in the accumulated biomass of about 40% compared to wild type. Peptidolytic activity studies showed no sign of compensatory mechanisms in the absence of AtPreP in mitochondria; in contrast we found a peptidolytic activity in the chloroplast membranes not related to AtPreP.

In addition to zinc located in the catalytic site, crystallographic data revealed two Mg-binding sites in the AtPreP structure. To further investigate the role of these Mg-binding sites, we have made AtPreP variants that are unable to bind metal ions. Our data shows that one of these sites located close to the catalytic site is important for the activity of AtPreP.

We also measured proteolytic activity of four human PreP-SNP variants and observed that the activity of all the hPreP-SNPs variants was lower; especially the hPreP-SNP (A525D) variant that displayed only 20-30 % of wild type activity. Interestingly, the activity was fully restored for all SNP-variants by addition of Mg2+.

5

Publications

The thesis is based on the following publications

I Nilsson S*, Bäckman HG*, Pesaresi P, Leister D, Glaser E. Deletion of an organellar peptidasome PreP affects early development in Arabidopsis thaliana. Plant Molecular Biology (2009) 71:497-508.

II Bäckman HG, Pessoa J, Eneqvist T, Glaser E. Binding of divalent cations is essential for the activity of the organellar peptidasome in Arabidopsis thaliana, AtPreP. FEBS Letters (2009) 583:2727-2733.

III Moreira Pinho C*, Björk Behnosh F*, Alikhani N, Bäckman HG, Eneqvist T, Fratiglioni L, Glaser E, Graff C. Genetic and biochemical studies of SNPs of the mitochondrial Aβ-degrading protease, hPreP. Neuroscience Letters (2010) 469:204- 208.

* These authors contributed equally

6

Abbreviations

IE Inner envelope membrane of chloroplast

IES Interenvelope space

IM Inner membrane

IMS Intermembrane space

MA Mitochondrial matrix

OE Outer envelope membrane of chloroplast

OM Outer membrane

PAM Presequence Associated Motor

PreP Presequence protease

PSI Photosystem 1

PSII Photosystem 2

TIC Translocase of the inner chloroplast envelope

TIM Translocase of the inner mitochondrial membrane

TM Transmembrane

TOC Translocase of the outer chloroplast envelope

TOM Translocase of the outer mitochondrial membrane

7

Mitochondria and chloroplasts

The mitochondrion is the power generating organelle that resides in the eukaryotic . It is generating adenosine triphosphate (ATP), which in turn can be used by the cell to do chemical work. Mitochondria are also involved in other tasks such as signaling, cellular differentiation, cell death, cell cycle control and cell growth [1]. The word mitochondrion is of Greek origin, where “mitos” stands for a thread and “chondrion” stands for a grain. In the year 1949 mitochondria were identified as the site of oxidative energy metabolism by Kennedy and Lehninger [2].

Chloroplasts are members of the plant organell class called plastids that all originate from protoplastids. The protoplastids differentiate during the plant development to form three major groups of plastids, the green chloroplasts, the coloured chromoplasts and the colourless leucoplasts. The plastids that are most abundant and most important are the chloroplasts. The word chloroplast is of Greek origin, where “chloro” stands for green and “plast” stands for form or entity. Chloroplasts harvest the energy of the sun to do work in form of splitting water and fixing carbon dioxide to produce sugars.

Origin and evolution

In 1970 Lynn Margulis proposed the endosymbiotic theory [3]. This theory was at first controversial, but today it is the most accepted theory about the origin of the intralcellular compartments that we now call mitochondria and chloroplasts. According to the endosymbiotic theory mitochondria arose from aerobic prokaryotes that were engulfed by and began to live in symbiosis with a primitive, anaerobic eukaryotic cell. The genomic data point towards α-proteobacteria as being the first cell to be engulfed. There are several important factors that support this theory. First of all, mitochondria are not synthesized by the cell, instead they originate from pre-existing organelles and are proliferated by a fission process. Secondly, mitochondria have its own genome and a transcription and translation system that is remnant to what is found in bacteria. Thirdly, there is a strong group correlation between found in mitochondria and those found in α-proteobacteria [4]. Finally, the inner membrane (IM) of mitochondria contains a certain lipid, such as cardiolipin, that only can be found in the bacterial plasma membrane and in the IM of mitochondria. The closest now living ancestor to mitochondria is Rickettsia prowazekii, which is an obligate intracellular parasite [5, 6].

Just like mitochondria, the evolution of the chloroplast is also the result of a single endosymbiotic event. More than one billion years ago a single eukaryotic cell containing mitochondria, engulfed and established an endosymbiotic relationship with a cyanobacterium. Around 450 million years ago these

8 cells colonized the planet outside the oceans, and by that time the engulfed cyanobacteria already had turned into what we know as chloroplasts [7].

The mitochondrial genome

Mitochondria harbor their own genome in form of a circular DNA. Mitochondrial genomes (mtDNA) have changed over time since the ancient endosymbiotic event. During evolution of mitochondria some genes have been lost or transfered from mitochondria to the cell nucleus resulting in a very small mtDNA. At the same time the organelle has retained much of its prokaryotic biochemistry, mtDNA now only encodes a small fraction ranging from 3 to 67 of the organelle’s proteins [8]. A free living α- proteobacteria such as Caulobacter contains over 3600 protein encoding genes. There are at least two distinct modes of how genetic loss has reduced the coding capacities of mitochondrial genomes, going from approximately 1600 proteins in the free-living ancestor to 67 or less in modern mitochondria. First, mitochondria have lost nonessential coding sequences, and rather than synthesize compounds such as amino acids, nucleosides and pyruvate, mitochondria now relay on import of these compounds. Secondly, mitochondria have lost essential genes due to extensive gene transfer to the nuclear genome, and therefore relay on import of these proteins. This transfer event has led to that nearly all of the genes required for the normal function and also propagation of mitochondria normally reside in the host nucleus [9, 10]. When comparing over 100 different mitochondrial genomes, the majority contains the same 13 genes encoding proteins of the respiratory chain and ATPsynthase, namely the cytochrome c oxidase subunit I, II and III, the ATP synthase subunits 6 and 8, the cytochrome b apoenzyme and the NADH dehydrogenase subunits 1-6 and 4L [11, 12].

Function and structure of mitochondria

Mitochondria can be found in almost all eukaryotic cells and are typically 0.5-1.5 µm wide and 1-2 µm long. They can vary in size, shape and number, especially when compared between different species. Mitochondria are organelles surrounded by two membranes, the outer mitochondrial membrane (OM) and the inner mitochondrial membrane (IM). The two membranes divide the organelle into two distinct compartments, the intermembrane space (IMS) and the mitochondrial matrix (MA). The composition of the OM is relatively simple, consisting of a phospholipid bilayer and proteins. The most abundant protein in the OM is porin, which makes the OM permeable to molecules of about 8 kDa or less. The inner membrane is more complex and contains the of the respiratory chain and oxidative phosphorylation as well as other proteins. The inner membrane is also highly convoluted that increases the surface area, forming folds called cristae. Cristae are connected to the IM by narrow tubular segments, called cristae junctions.

9

Mitochondria play an important role in metabolism, apoptosis, disease and aging. Mitochondria are also the site of oxidative phosphorylation, which is essential for the production of ATP. The energy required for the ATP production predominantly comes from the oxidation of reduced compounds such as NADH and FADH2, which are generated in the citric acid cycle and during β-oxidation of fatty acids. The oxidation of NADH and FADH2 releases electrons that are funneled to oxygen through a chain of electron carriers in the respiratory chain. This electron transfer releases energy that is used to transfer protons across the IM from the matrix to the IMS side. This proton translocation is generates an electrochemical proton motive force (PMF). The PMF consists of a membrane potential (∆ѱ) and a trans-membrane proton gradient (∆pH) [13]. The potential energy trapped in the PMF can then be used to synthesize ATP from ADP and Pi by the ATP synthase located in the IM of mitochondria. ATP is then exported from the mitochondria out to different parts of the cell where it provides the energy required for a vast type of reactions.

The chloroplast genome

Chloroplasts harbor their own genome in form of the circular DNA and just like the mitochondrial genome, the chloroplast genome is also of procaryotic origin. During the evolution of the plastids, many genes have been transferred to the cell nucleus. Compared to cyanobacterial genomes, which contain between 3168 (Synechocystis sp.) and 7400 (Notoc punctiforme) protein encoding genes, the chloroplast genome has been substantially reduced in size and vary between 58 (Euglena gracilis) and 200 protein coding genes (Porphyra purpurea). In chloroplasts of green plants there are about 100 protein encoding genes [14, 15].

Function and structure of chloroplasts

The chloroplasts are disc shaped organelles that belong to a broader group of plastids. Chloroplasts are usually between 5-10 µm in diameter and about 1 µm thick, and are surrounded by two membranes, the outer envelope (OE) membrane and the inner envelope (IE) membrane. The OE membrane is permeable to small molecules up to about 10 kDa in size, but it is impermeable to bigger molecules such as proteins. The inner envelope membrane works as a permeability barrier, and molecules must be transported across the membrane via specific translocators. The compartment between these two envelopes is called the interenvelope space (IES). Enclosed by the inner envelope membrane is the chloroplast stroma. The stroma contains a third membrane system, called the thylakoids. The thylakoids consist of a network of small disks that is forming stacks or grana. The grana stacks are in turn connected by a non-stacked membrane system called stroma lamellae. The thylakoid membrane is one continuous membrane system enclosing the thylakoid lumen [7].

10

Chloroplasts have a very important function within plant cells, they constitute the location for photosynthesis and the fixation of CO2 along with other essential processes such as starch metabolism, biosynthesis of lipids and secondary metabolites. The main components of the thylakoids are the two photosystems (PSI and PSII), the cytochrome b6f complex and the ATP synthase. In the photosystems the light is harvested and the energy is used for creating a charge separation within PSII. This then leads to the oxidation of water to oxygen in the thylakoid lumen, and four electrons and two protons are released in this process. The electrons are then transported via the Q cycle and b6f complex and further to the PSI and causing more protons to be transferred into the lumen. The final electron acceptor is NADP+ that is reduced to NADPH. The protons that are transferred to the thylakoid lumen generate a proton gradient that is used to generate ATP by the ATP synthase. The energy stored in form of ATP can then be used in the fixation of CO2, an important step in the plant metabolism [16].

Protein targeting and import into the mitochondrial matrix

Since the vast majority of the protein coding genes has been transferred to the cell nucleus, mitochondria have to rely on protein import for its normal function. The genes are transcribed in the cell nucleus and translated to precursor proteins on cytosolic ribosomes and then posttranslationally transferred to its final destination in mitochondria. From the cytosol the precursor proteins are translocated through the outer mitochondrial membrane via the TOM (Translocase of the outer mitochondrial membrane) complex and then through the inner membrane via the TIM (Translocase of the inner mitochondrial membrane) complex. Finally, in the mitochondrial matrix the protein is further processed to reach a final and functional conformation (for recent review see [17]).

Protein targeting to mitochondria relies on two main types of targeting signals. The first and most common way of targeting mitochondrial proteins is by the use of cleavable amino-terminal extensions in the mitochondrial precursor proteins, called presequences. About 70% of mitochondrial proteins are recognized via an N-terminal presequence [18]. Presequences vary in length and range in size between 18 to 136 amino acid residues in plants with an average length of 42 residues; most presequences are in the range of 20-60 amino acid residues long [19]. The presequences share little sequence similarity except that they are generally enriched in positively charged, hydroxylated and hydrophobic amino acid residues, while acidic residues are almost absent [19, 20]. Further, it has been shown that presequences have the ability to form amphiphilic α-helices in a membrane mimetic environment [21- 25]. The presequence is cleaved off by the mitochondrial processing peptidase (MPP) [26]. The second type of signals consists of an internal targeting sequence located internally in the mature protein. This signal is not cleaved off after import into mitochondria. In the cytosol, precursor proteins are protected by molecular chaperones, which maintain precursor proteins in an unfolded import competent conformation [27].

11

With the aid of chaperones and other cytosolic factors, precursor proteins are recognized by receptors on the mitochondrial surface. There are three main receptors for recognition of precursor proteins on the outer membrane of mitochondria, Tom20, Tom22 (in plants Tom9) and Tom70 (absent in plants). Tom70 binds proteins with internal non-cleavable targeting signals [28, 29]. The main receptor for N- terminal presequences is Tom20, which binds presequences via hydrophobic interactions in a hydrophobic grove. Studies have shown that A. thaliana Tom20 is a non-essential protein [30]. Tom20 redirects the bound precursor protein to Tom40 to begin the translocation [27, 31].

Tom40 is a β-barrel protein and acts as the translocation pore for almost all mitochondrial proteins. Tom40 was the first mitochondrial membrane protein that was shown to be essential for yeast viability [32]. Tom40 has also been shown to be essential in Arabidopsis [33]. The precursor protein is then assisted by chaperones (Tim8-10 and 13) in the IMS, where it interacts with translocase of the mitochondrial IM [17].

Many proteins that are destined for IMS have a cysteine motif (or MISS signal) as a targeting signal that forms an intramolecular disulfide bond, which in turn traps the protein in the IMS. Mia40 and Erv1 form a disulfide bond relay system where Mia40 oxidizes substrates and Erv1 is reoxidizing Mia40 [17, 27, 34-36].

In the mitochondrial IM there are at least three complexes that are responsible for the translocation across the IM to the MA and insertion of proteins into the IM. The two major complexes are the TIM23 and TIM22 complexes. The TIM23 complex is responsible for the translocation of all precursors of matrix proteins and most of the IM proteins and many IMS proteins. The TIM22 complex is responsible for the translocation and IM insertion of carrier type membrane proteins containing internal targeting signals [37]. The third complex is OXA1 that is a component involved in the insertion of MA imported or mitochondria encoded proteins into the IM [38]. Compared to the outer membrane import components, the Arabidopsis inner membrane components are more similar to the ones found in yeast [39, 40].

The initial translocation through the TIM23 complex is dependent on an electrical membrane potential (∆ѱ). The TIM23 complex consists of two structurally and functionally distinct forms [34]. The initial form, termed TIM23-SORT, contains the channel forming proteins Tim23 and Tim17, together with its associated proteins Tim50 and Tim21. All these components of the TIM23-SORT complex are fundamental for the translocation of precursor proteins from the Tom40 to the matrix or for insertion into the inner membrane [41]. In yeast the second form of the TIM23 complex is called the TIM23- PAM complex. This complex also contains Tim17, Tim23 and Tim50 along with the Presequence translocase Associated Motor (PAM) complex, comprising mtHSP70 (mitochondrial heat shock protein 70), Tim44 and its associated co-chaperones; Pam16, Pam17 and Pam18 [42, 43]. The TIM23- PAM complex drives the translocation through the inner membrane via ATPase activity and cyclic

12 binding of the presequence to promote transfer of the precursor protein towards the matrix [41]. Pam16, 17 and 18 are molecular chaperones that are thought to modulate mtHSP70 (mitochondrial heat shock protein, 70 kDa) activity and association with the TIM23 channel [44, 45]. With the exception of Pam17, which is absent in plants, there is a high level of similarity between the yeast inner membrane proteins and the ones found in A. thaliana; therefore it is likely that these pathways are also functionally conserved in plants.

Mitochondrial processing peptidases

When the mitochondrial precursor proteins have been transported across the inner membrane into the matrix, there are processing steps that are necessary before the imported protein is fully functional. There are several proteins involved in these processing steps, the mitochondrial processing peptidase (MPP), the inner membrane peptidase (IMP), the Octapeptidyl aminopeptidase 1 (Oct1), and the intermediate cleaving peptidase 55 (Icp55) [46, 47].

MPP is a ~ 100 kDa heterodimeric encoded by two genes, α-MPP and β-MPP, where both subunits are essential and highly conserved [46]. MPP cleaves off the N-terminal presequence from the translocated precursor protein. The cleavage sites typically contain an arginine at position -2 from the scissile peptide bond and an aromatic residue at position +1 [48]. The α-MPP is responsible for the peptide binding and the β-MPP contains the for proteolysis. MPP belongs to the M16C family in the MEROPS protease database [49]. The β-MPP contains an essential inverted zinc binding motif (HXXEH), thereby placing it among the metalloproteases [50]. Although there is similarity (31 and 41% for the α- and β-MPP, respectively) between the Arabidopsis and the yeast counterparts, there are significant differences in these organisms. In Saccharomyces cerevisiae the MPP resides in the MA as a soluble protease [51], whereas in plants it is a part of the cytochrome bc1 complex of the electron transport chain in the IM. This means that the plant MPP is bi-functional, and is involved in both processing and electron transfer [52-55]. In 2001 the 3D structure of MPP was solved, both for the active form and an inactive variant. The structures revealed a large central cavity containing the active site situated between the α- and β-MPP. This cavity is lined with hydrophilic amino acids, including many glutamate and aspartate residues, from both subunits. This negative charge could facilitate the recognition and binding of the basic presequences of precursor proteins [48].

IMP is a serine endopeptidase that catalyzes the maturation of mitochondrial precursor proteins delivered to the intermembrane space. In S. cerevisiae, the exists as a heterodimeric complex composed of two different subunits: Imp1 and Imp2, and the auxiliary protein Som1. Imp1 and Imp2

13 are both bound to the outer face of the IM through an N-terminal membrane spanning domain and have a C-terminal catalytic site facing the IMS [47, 56].

Oct1 is a soluble monomeric metalloprotease in the mitochondrial matrix. The protease has a putative zink-binding motif H-E-X-X-H that is required for the catalytic activity, and two highly conserved cysteine residues are important for the stability of the protein. Oct1 only recognizes and cleaves proteins that have already been processed by MPP. The protease was formerly known as MIP (mitochondrial intermediate peptidase), and it removes an additional octapeptide from the precursor after MPP cleavage. The octapeptide is required for an efficient first processing by MPP, and it is believed that the function of the octapeptide might be to provide a more compatible MPP cleavage site by separating the first processing site from a positive structure in the mature part of the protein. Finally it is also believed that the removal of the octapeptide is contributing to a more stable mature protein [47, 57, 58].

Icp55 is a novel metalloprotease belonging to the aminopeptidase P family and is found in the mitochondrial matrix. In contrast to other aminopeptidase P proteases it does not require a proline residue at the cleavage site, which gives the protease a different specificity compared to other proteases in this family. Instead the cleavage site motif for Icp55 is: R-X-(F/L/Y) | (S/A). Icp55 cleaves off a single phenylalanine, tyrosine or leucine residues appearing at the N-terminus of precursor import intermediates after initial processing by MPP. These three amino acids are classified as destabilizing amino acids according to the N-end rule [18, 47, 59].

Protein targeting and import into chloroplast stroma

As mitochondria, chloroplasts are dependent on protein import from the cytosol. The precursor proteins are translocated through the outer chloroplast envelope via the TOC (Translocase of the outer chloroplast envelope) complex and then through the inner envelope via the TIC (Translocase of the inner chloroplast envelope) complex. In the chloroplast stroma the protein is further processed to reach its final and functional conformation.

The precursor protein carries an amino-terminal extension called a transit peptide. Transit peptides are remarkably diverse in both length and sequence, as they vary from 20 to >100 residues, and possess no extended blocks of sequence conservation. They seem to have some conserved properties in form of abundance of hydroxylated residues (serine in particular), and they are deficient in acidic residues giving them a net positive charge; so in this regard transit peptides are very similar to mitochondrial presequences [60].

14

Hsp70 is one of the cytosolic factors that is believed to play a role in the cytosol until the precursor protein reaches the import machinery of the chloroplast outer surface. Most of the chloroplast transit peptides have indeed an Hsp70 and direct contact between Hsp70 and transit peptides has been reported, but it is still unclear how important this interaction is for the import of a specific precursor protein [61].

Other cytosolic factors are found in the 14-3-3 that includes regulatory molecules and chaperones that bind specifically to phosphorylated proteins and thereby mediate various signal transduction processes and also protein translocation. Many chloroplast transit peptides contain a 14- 3-3-binding phosphopeptide motif, and it has been shown that 14-3-3 proteins can form a type of “guidance complex” together with Hsp70 and the precursor protein, directing precursors to the surface of chloroplasts [60, 62].

The TOC complex is responsible for the precursor protein recognition at the chloroplast surface and for outer envelope translocation. The TOC core complex is composed of three different proteins, Toc159, Toc34 and Toc75, according to their predicted molecular masses. These three proteins constitute a complex that is between 500 and 1000 kDa in size, as there are reports of different possible stoichiometric configurations.

The Toc159 and Toc34 proteins can be regarded as receptors as they control precursor protein recognition. They are both GTPases and are anchored to the membrane via the C-terminal domain. The third component of the core complex, Toc75 is embedded within the outer membrane, is similar to the functionally equivalent Tom40 protein of mitochondria and has a β-barrel structure [63, 64]. Toc75 has a topology of 16-18 transmembrane (TM) strands that together form an aqueous pore with a minimal diameter of approximately 14 Å. This is very similar in size to the Tom40 channel, as well as other protein translocation pores, and it is sufficient to accept only largely unfolded substrates [65].

After the GTP dependent translocation through the Toc75, the precursor protein is translocated through the TIC complex in the inner envelope to finally reach the chloroplast stroma. The configuration and function of the TIC complex is still under debate, but several actual or putative components have been identified – Tic110, Tic62, Tic55, Tic40, Tic32, Tic22 and Tic20. It is believed that Toc12, Hsp70 and Tic22 act to facilitate the passage of precursor proteins across the interenvelope space. Tic110 has been proposed to be a channel forming protein consisting of six transmembrane domains and a scaffold with two N-terminal transmembrane domains followed by a large soluble domain for binding of transit peptides and other stromal translocon components [66]. The translocation through the Tim110 channel occurs in an ATP-dependent manner, where a C-terminal domain of Tic40 stimulates the ATPase activity of Hsp93, which in turn facilitates the import of the precursor protein to the chloroplast stroma [60].

15

Stromal processing peptidase

When the precursor protein has been translocated into the chloroplast stroma it is processed by the stromal processing protease (SPP). Studies in A. thaliana using T-DNA insertion lines have shown that SPP is an essential protease [67].

SPP is a soluble metalloendopeptidase in the chloroplast stroma. It contains an inverted Zn-binding motif, HXXEH, and belongs to the (MEROPS) metallopeptidase family M16B. It recognizes and binds to transit peptides of precursor proteins and cleaves off the transit peptide in one endoproteolytic step, which leads to the formation of a functional mature protein. After cleavage, the transit peptide is still bound to SPP and can then be trimmed [68]. The free transit peptide may then be degraded by the Presequence Protease, PreP.

Dual-targeting to mitochondria and chloroplasts

Despite that the import mechanisms of mitochondria and chloroplast have been developed independently there are proteins that are imported to each organelle using the same targeting sequence, these proteins are called dual-targeted proteins. More than 100 proteins have been reported to be dual- targeted, usually determined by experimental procedures using fluorescent markers such as GFP, followed by specific import assays to each organelle to establish that a single protein is a substrate for both import machineries. It appears that the majority of the dual-targeted proteins are using an identical targeting peptide, which in turn is recognized by receptors on both organelles [69]. These targeting peptides are called ‘ambiguous’, as they carry a dual function [70]. Most of the dual-targeted proteins are soluble matrix/stromal proteins involved in processes such as DNA replication, transcription and translation, protein processing and other metabolic processes. Only a few are membrane-bound, and most of them are bound to the outer membrane [71].

16

Proteolysis in mitochondria and chloroplasts

The ATP dependent proteases

The ATP dependent proteases are a group of proteases that are dependent on ATP to complete their function in the cell. ATP is not needed for the catalytic function of the protesase, instead it is used to unfold larger proteins to make them accessible for degradation by the protease. Proteases usually have the catalytic site hidden in a cavity to prevent unspecific proteolysis. The major ATP dependent proteases in mitochondria and chloroplasts are ClpP, FtsH and Lon [72].

Clp proteases

The Clp endopeptidases are ATP dependent serine proteases belonging to the S14 Merops family of proteases, which are involved in the degradation of misfolded proteins. There are 26 genes encoding Clp proteases in A. thaliana. Clp proteases are multi-subunit enzymes, and the ATPase domain and the catalytic domain are residing in different subunits. In structural terms they are very similar to the proteasome 26S in eukaryotes. Most of the knowledge about Clp proteases comes from investigations in E. coli. In E. coli the protease is composed by two ClpP subunits, which form a barrel structure with narrow (~10 Å) axial openings, with the catalytic sites facing the inner chamber. In addition to this proteolytic chamber there are ATP dependent chaperones, ClpA or ClpX (members of the HSP100 family of chaperones in E. coli) forming dynamic hexameric rings that interact with the core complex at the entrances, functioning as unfoldases and recognition elements [73, 74]. Plants have a high number of Clp proteins and in A. thaliana plastids there are at least 15 different Clp proteases. There are five serine-type ClpP (ClpP1, ClpP3–ClpP6) proteases; four ClpP-related ClpR (ClpR1–ClpR4) proteins; three Clp AAA+ chaperones (ClpC1, ClpC2 and ClpD), which are similar to ClpA from E. coli and three members (ClpS1, ClpS2, ClpT) with unknown functions. The ClpR proteins are homologous to the ClpP proteins, but are inactive as they lack the three conserved (Ser and His) catalytic amino-acid residues. The ClpS1 and ClpS2 are only found in land plants and have sequence homology to the amino-terminus of ClpC [75, 76].

FtsH proteases

FtsH proteases are membrane bound ATP dependent metalloproteases that harbour an AAA (or triple A) domain (ATPases associated with various cellular activities) [77]. The term FtsH comes from the E. coli enzyme (Filamentous temperature sensitive H) [78]. FtsH proteases are found in eubacteria as well as in mitochondria and chloroplasts. The main function of organellar FtsH proteases is selective degradation of non-assembled, incompletely assembled and damaged membrane-anchored proteins. It is believed that the thylakoid FtsH protease has a function in the degradation of the high turnover D1 17 protein of the PSII reaction center [79]. The FtsH protease is also involved in processing of precursor proteins and degradation of regulatory proteins. Substrate proteins are unfolded and translocated through the central pore of the ATPase domain into the proteolytic chamber, where the polypeptide chains are cleaved into short peptides.

In mitochondria there are two types of AAA protease complexes, which differ in their topology in the inner membrane; there are i-AAA proteases that are active in the intermembrane space and m-AAA proteases with catalytic sites exposed to the mitochondrial matrix. These proteases have homologous subunits that are assembled into homo- or hetero-oligomeric complexes, forming a hexameric ring. FtsH protease subunits have two trans-membrane domains in the N-terminal portion, followed by an ATPase domain, and a zinc-binding motif, which serves as the catalytic site of the protease. Most eubacteria have only one gene encoding FtsH, whereas three genes have been found in yeast and humans, while 17 orthologs have been identified in A. thaliana [72, 80]. 5 of these 17 genes are coding for inactive isoforms (FtsHi 1-5) targeted to chloroplasts, and are all lacking the zinc-binding motif (HEXXH) [81]. Out of the 12 active isoforms three (AtFtsH3, 4 and 10) are located in mitochondria and 8 (AtFtsH1, 2, 5-10, 12) in chloroplasts. FtsH11 is dually targeted to both mitochondria and chloroplasts [82].

Lon proteases

Lon proteases are members of the S16 family in the MEROPS protease database. They are serine proteases that are responsible for the degradation of abnormal, damaged and unstable proteins. Lon proteases have no membrane-spanning domain and contain the AAA and protease domains in one polypeptide. In contrast to the Ser-His-Asp triad at the catalytic site, Lon proteases consist of a Ser- Lys dyad [83, 84]. A crystal structure of Lon in E. coli has been determined and shows that Lon forms a hexameric ring [85]. Lon proteases have been reported as mitochondrial proteases, while some studies have also predicted their presence in chloroplasts and peroxisomes [86, 87]. In E. coli Lon is divided into two subfamilies, LonA and LonB. Both contain an ATPase domain as the proteolytic domain, but differ in that the LonA enzymes contain a large N-terminal domain, whereas LonB lacks this domain. Instead LonB harbours an N-terminal transmembrane domain. In A. thaliana there are 11 genes encoding Lon family proteases, named AtLon1-11, however there are data suggesting that only four Lon proteases are functional in organellar proteolysis in A. thaliana (AtLon1-4) [72, 88].

18

The ATP independent proteases

Rhomboid proteases

The name ‘rhomboid’ comes from a discovery made in the field of Drosophila genetics. When analysing the phenotype of a mutant larval cuticle, a misshapen rhombus-like head skeleton of the mutant embryo was discovered, which gave rise to the name rhomboid protease [89]. The rhomboid gene was identified in Drosophila screens of the late 1970s and early 1980s [90], and it was cloned and sequenced by Bier and colleagues in 1990. They discovered a protein with no homology to any known sequence at the time [91]. Sequencing of genomes from diverse organisms has shown that rhomboids indeed are present in every form of cellular life [92]. There are three different topological forms of rhomboid proteases. The simplest consist of a 6 TM core, which itself is the smallest catalytically active unit. This form is common in bacteria, but can also be found rarely in eukaryotes including . The second form has the 6 TM core plus an additional TM segment (6+1 TM form). This form is represented by five out of the seven Drosophila, human and mouse rhomboid proteins. The last form is the 7 TM rhomboid proteases that exist in endosymbiotic organelles, which adds another TM domain to the 6+1 TM form. The Rhomboid-1 protease in Drosophila is localized in the where it cleaves a substrate called Spitz. The released product is secreted to activate epidermal growth factor (EGF) signaling in neighboring cells.

Rhomboid proteases are serine proteases belonging to the S54 family of the MEROPS peptidase database, and contain the characteristic Ser-His catalytic dyad. In mitochondria rhomboids are intramembrane proteases consisting of seven transmembrane helices [89]. The fourth transmembrane segment is slightly tilted in relation to the others, and enters the center of the protein as an extended loop, converting to an α-helix at the catalytic serine. There are 15 annotated rhomboid proteases in A. thaliana. Two of these (AtRbl1 and 2) have been located to the Gogi apparatus [93], the subcellular localization of most of the others is predicted to be in mitochondria. The AtRbl9 and 10 are predicted to be localized in chloroplasts using the programs TargetP and Predator. There are no known substrates for the Arabidopsis rhomboid proteases; hence their function is not well documented in plants.

Deg proteases

Degradation of periplasmic proteins DegP proteases [94], also referred to as high temperature requirement A (HtrA) proteases was initially characterized in E. coli in 1989 [95, 96]. Since then DegP/HtrA proteases have been characterized as ATP independent temperature sensitive serine endopeptidases and have been found in almost every organism, where the ones in metazoans are designated as ‘HtrA proteases’ and the ones in plants as ‘DegP proteases’. The Deg proteases belong

19 to the S1B subfamily in the MEROPS protease database, with a comprised of His, Asp and Ser [97, 98].

Deg proteases form homo-oligomeric complexes and contain a trimer as the basic structural unit, formed by the interactions of protease domains. These trimers may in turn assemble into higher oligomeric complexes, such as hexamers that have been reported from studies in E. coli [97], and Deg1 from A. thaliana [99], or even 12-mers and 24-mers, as described for E. coli DegP [100]. There are 16 genes coding for Deg proteases in A. thaliana AtDeg1-16. AtDeg1, 2, 5 and 8 have all been localized to the chloroplast. AtDeg2 is localized on the stromal side of the thylakoid membrane, while AtDeg1, 5 and 8 are located in the thylakoid lumen. AtDeg3, 4, 6, 10-13 and 15 have all been predicted to be mitochondrial [72]. AtDeg1 has been shown to be an essential protease; even reduced level of AtDeg1 gave rise to a relatively strong phenotype in form of reduced growth and early flowering in combination with a chlorotic phenotype [101]. Reported substrates for the AtDeg proteases are proteins of the PSII, such as the high turnover protein D1, and other proteins of the PSII as the D2, CP43 and CP47 [94].

OOP

Organellar oligopeptidase (OOP) is a peptide-degrading metalloprotease belonging to the M3A family in the MEROPS database. OOP is a soluble protein, dually localized to the mitochondrial matrix and chloroplast stroma. The recently solved structure of OOP revealed an ellipsoidal shape consisting of two major domains enclosing the catalytic cavity of 3000 Å3. The protease degrades short peptides with a size restriction from 8 to 23 amino acid residues. Experiments have shown that OOP is able to cleave both N- and C-terminal fragments of mitochondrial presequences, as well as short transit peptides and other short peptides [102].

PreP

Presequence protease called PreP is an ATP independent protease that is located in both mitochondria and chloroplasts. PreP will be described in more detail below as it is the main subject of this thesis.

20

Material & Methods

Paper I

Generation of mutant Arabidopsis plants, single and double knockouts

Several T-DNA insertion lines for both AtPreP1 and AtPreP2 were ordered from the Nottingham Arabidopsis Stock Centre and the Arabidopsis Biological Resource Centre. The samples were genotyped using PCR in combination with Western blot analysis. Homozygous double knockout plants were generated by crossing of homozygous single mutants. The double mutant plants were identified using PCR screening.

Growth conditions of A. thaliana plants

Plants were grown on MSO (Murashige and Skoog medium) agar plates for genotyping in a growth chamber under continuous light. For all other experiments, plants were grown on soil in a growth chamber with either 8 or 16 hours of light each day, with a night and day temperature of 18 °C and 22 °C, respectively.

Purification of mitochondria and chloroplasts

To obtain a crude mitochondrial fraction, 5 g of leaves were homogenized with sand and grinding buffer. The grinded leaves were then filtrated and the homogenate was centrifuged and washed [103]. The pellet containing crude mitochondria was resuspended in wash buffer and used for respiratory measurements.

Highly purified mitochondria were purified using the methods described in [103].

To isolate mitochondrial matrix and membrane fractions, highly purified mitochondria fractions were incubated in a degradation buffer to obtain osmotic swelling, followed by sonication and centrifugation. The supernatant was used as mitochondrial matrix fraction and the pellet was further washed in degradation buffer and sonicated and centrifuged one more time. The resulting pellet was then resuspended in degradation buffer and was used as membrane fraction.

Chloroplasts were isolated by using 10 g of tissue that was mixed in a blender in homogenization buffer. The resulting homogenate was filtered and centrifuged, and then loaded onto a Percoll gradient and centrifuged. Chloroplasts were then harvested from the gradient for further use [104].

21

Stroma and chloroplast membrane isolation was performed by osmotic swelling of chloroplasts in degradation buffer followed by sonication and centrifugation steps. The supernatant was used as stroma fraction and the pellet was further washed and resuspended in degradation buffer and used as membrane fraction.

Thylakoids were isolated using 2 g of leaves, which were homogeniezed in a preparation buffer using a mixer. The homogenate was then filtered and centrifuged and the pellet was washed and resuspended in lysis buffer and centrifuged. The resulting pellet was resuspended in incubation buffer and homogenized. Thylakoid membranes were collected after centrifugation.

Degradation assays

Degradation assays were performed using a HEPES degradation buffer at pH 8.0. For the analysis of the mitochondrial and chloroplast membrane fractions 0.1 µg of the pF1β(2-54) peptide was degraded for two hours at 30 °C by 2,5 µg of membrane protein. Samples were separated on 15% SDS-PAGE and transferred by Western blot to Hybond ECL membranes and treated with primary and secondary antibodies for detection.

For the mitochondrial matrix and chloroplast stroma fractions 1 µg of pF1β(2-54) peptide was degraded for two hours at 30 °C. Samples were separated on 15% SDS-PAGE and stained with Coomassie blue.

Respiratory measurements

Respiratory measurements were performed in similar fashion as described in [103].

Protein gel and immunoblot analysis of thylakoid membranes

2 mg chlorophyll per ml sample was separated on 15 % SDS-PAGE followed by Blue Native PAGE. Quantification of results was done using scanned triplicates, which were analyzed using the PDQuest software. Immunoblot analysis was done using a semi dry Western blot transfer followed by detection with antibodies for Lhcb2 and the F1β subunit of the ATPase.

Protein extractions and analysis of AtPreP1 and AtPreP2 content

Plant tissues from mature plants were frozen and grinded. The resulting tissue powder was resuspended in a Tris-buffer and centrifuged. 10 µg of protein was loaded onto a 12 % SDS-PAGE

22 and PreP content was analyzed using Western blot with an antibody recognizing both AtPreP1 and AtPreP2.

Biomass

Plants were cut at the base excluding the roots. Harvested plants were kept dry in the dark for more than 3 days to fully dry before they were weighted.

Paper II

Generation of AtPreP1 variants

To generate the AtPreP variants mutations were introduced at D705N, D859N, D861N for the inner Mg-binding site and D828N, T947N, D952N for the outer Mg-binding site. All mutations were introduced to wild type AtPreP1 gene on pGEX-6P-2 plasmid (Amersham Biosciences) using the Quick-Change Site Directed Mutagenesis Kit (Stratagene). All constructs were verified by DNA sequencing.

Expression and purification of AtPreP1 and its variants

Wild type AtPreP and the three variants were linked to a Glutathione S- (GST) tag at the N-terminus. All constructs were transformed into E. coli BL21 (DE3). Colonies were inoculated in LB medium and overexpressed with 1mM Isopropyl β-D-1-thiogalactopyranoside (IPTG) and incubated at

30 °C to OD600 of 2. Cells were harvested by centrifugation and stored at -80 °C. The pellet was resuspended in phosphate-buffered saline buffer and cells were lysed with lysozyme and sonication. The soluble protein fraction was separated by centrifugation and filtration and incubated with Glutathione Sepharose® beads (Amersham Biosciences) in phosphate-buffered saline buffer at 4 °C for 4 hours. Bead solution was washed and bound AtPreP was cleaved over night using PreScission™ Protease. After cleavage the solution was centrifuged and AtPreP was collected from the supernatant. Protein samples were analyzed on 12 % SDS-PAGE, and final protein concentration was determined using BioRad Protein Assay (BioRad). The protein was stored at -20 °C.

23

Degradation assays

Degradation assays with hPrPss(1-28), Cecropin A(1-33), Penetratin, Galanin and the F1β presequence peptides were done using 1 µg of peptide and 1 µg of AtPreP incubated in HEPES buffer at 30 °C for 1 hour. Degradation was analyzed on 15 % SDS-PAGE. Degradation of the 11 amino acids C1 peptide was done using 1 µg of peptide and 1 µg of AtPreP incubated in HEPES buffer at 30 °C for 1 hour and analyzed by electrophoresis on 1 % agarose gels and the C1 peptide was detected by UV light.

Substrate V (R&D Systems), a bradykinin derived peptide of 9 amino acids was used to monitor the kinetics of proteolysis using a fluorometer. For these experiments 1 µg of Substrate V and 1 µg of AtPreP protein were mixed with a HEPES degradation buffer in a cuvette. Degradation data for Substrate V was recorded for 90 seconds with excitation and emission wavelength set at 320 nm and 405 nm, respectively.

Paper III

Generation of hPreP-SNP variants

To generate the hPreP-SNP variants substitutions were introduced at L116V, F140S, A525D and I924M. All mutations were introduced to the wild type AtPreP1 gene in the pGEX-6P-2 plasmid (Amersham Biosciences) using the Quick-Change Site Directed Mutagenesis Kit (Stratagene). All constructs were verified by DNA sequencing.

Degradation assays

Degradation assays using hPreP and hPreP-SNP variants were done using 1 µg of hPreP protein and 1

µg of following peptide substrates; Aβ(1-40) or Aβ(1-42), or pF1β(2-54). hPreP and the substrate were mixed in a HEPES degradation buffer and incubated at 37 °C for 5-30 minutes (for pF1β(2-54) and for 30 minutes for Aβ(1-40) or Aβ(1-42). Degradation was analyzed on 10-20 % Tris-Tricine gels (Bio-Rad).

Substrate V was used to monitor the kinetics of proteolysis using a fluorometer. For these experiments 1,5 µg of Substrate V and 1 µg of hPreP protein were mixed with a HEPES degradation buffer in a cuvette. Degradation data for Substrate V was recorded for 120 seconds.

24

Summary of results

The Presequence protease, PreP

After import into the respective organelle the targeting peptide is cleaved off. In mitochondria the presequence is cleaved off by the mitochondrial processing peptidase (MPP) and in chloroplasts the transit peptide is cleaved off by the stromal processing peptidase (SPP). The mature proteins are then able to fold into the native form and fulfil various important activities inside the organelles. The free targeting peptides are subsequently degraded by the Presequence protease (PreP).

PreP is a 110 kDa zinc metallooligopeptidase that belongs to the M16C family of peptidases (MEROPS database) [49]. The M16 peptidases are commonly referred to as inverzincins, due to an inverted zinc-binding motif, HXXEH, at the active site [105]. In A. thaliana there are two isoforms of PreP, referred to as AtPreP1 and AtPreP2, with high sequence conservation (86% identity). AtPreP1 and AtPreP2 are dually targeted to both the mitochondrial matrix and the chloroplast stroma [106]. The human homolog of AtPreP, hPreP (or MP1 for metalloprotease 1) is located in the mitochondrial matrix as is the yeast homolog MOP112 (or CYM1p) [107].

The crystal structure of AtPreP1(E80Q) with a bound peptide in the active site was determined at 2.1 Å [108]. The structure showed that the 995-residue polypeptide folds into four topologically similar domains, which form two bowl-shaped halves connected by a V-shaped hinge that protrudes from the protease [108]. Although the inverted zinc binding motif of the active site (H77XXEH81) is positioned in the N-terminal domain, distal residues (Y854, R848) from the C-terminal domain come together to complete the active site. This leads to the hypothesis that the shift between an open and closed state is crucial for the activity of the protease [108]. Magnesium ions were added to facilitate the crystallization of AtPreP1 and the structure revealed two hydrated magnesium ions coordinated by acidic residues [108]. The structure of AtPreP1 revealed a unique proteolytic chamber of more than 10 000 Å3. Within this chamber lies the active site and it is estimated that the chamber has room for peptides up to about 65 amino acids in length. Mitochondrial and chloroplastic targeting peptides are about 30-50 amino acids long but do not show any sequence similarity. On the other hand they share some overall properties as they are generally enriched in positively charged, hydroxylated and hydrophobic amino acids [20]. Presequences also have the ability to form a positively charged amphipathic α-helix in the presence of a membrane-like environment [109]. The α-helix formation has also been observed for transit peptides [110]. Substrate specificity experiments have shown that PreP is actually a general peptidase and that it cleaves not just presequences and transit peptides, but also other unstructured peptides between 9-65 amino acids in length [106].

25

The inside surface of the chamber is negatively charged and it is proposed that the binding of a positively charged peptide induces the closing of the protease. After the chamber has been closed, the peptide can be cleaved and the protease is restored to an open conformation. This novel mechanism of action has led to that PreP and other M16 proteases now are referred to as peptidasomes [108, 111].

In general, proteolysis plays a vital role in the cell. Proteins and peptides have to be continuously degraded to sustain cell homeostasis. Presequences have membrane binding properties and also the ability to penetrate membranes, which can lead to loss of membrane potential and uncoupling of respiration [112]. Accumulation of free presequences could be toxic to the cell, and therefore they are proteolytically removed by PreP. It has been shown that the human homolog hPreP is able to degrade amyloid-beta peptide (Aβ) [113], the toxic agent that accumulates in mitochondria in Alzheimers disease [114].

When a mutation of a single nucleotide occurs within a genome and the frequency of this change is greater than 1 % it is called a SNP (single nucleotide polymorphism). SNPs can be either silent (synonymous), which do not lead to a change in the amino-acid composition, or non-synonymous that will give rise to a change in the amino-acid composition.

Our goal was to further study the AtPreP by constructing PreP knockout plants and make comparison with wild type plants, and to investigate the role of the magnesium-binding sites found in the AtPreP1 structure. Further, we localized four non-synonymous SNPs in hPreP that were situated in structural positions that might be important for the catalytic activity. We were interested in the role of these SNPs found in hPreP, and we compared the activity of the hPreP-SNP variants with wild type hPreP.

Paper I - Deletion of an organellar peptidasome PreP affects early development in Arabidopsis thaliana

After extensive investigation of the two PreP isoforms AtPreP1 and AtPreP2 in our laboratory we wanted to study the effect of PreP peptidasome deletion in A. thaliana. The aim of this project was to study the in vivo effects of PreP knockout and how the plant would cope with the toxic stress caused by a possible accumulation of peptides. The aim was to create single knockouts, atprep1 and atprep2, and double knockout atprep1 atprep2 strains.

T-DNA (Salk) insertion lines of A. thaliana (for both isoforms) was ordered and screened for homozygous single mutants. The homozygous single mutants of AtPreP1 and AtPreP2 were then crossed in order to generate a double knockout strain. According to the Genevestigator database,

26 which is based on micro array data, PreP1 is the most abundant isoform of the two isoforms and our Western blot analysis supported this data. Immunoblot analysis of different wild type plant tissues showed that PreP is most abundant in flowers and siliques, but also present at significant levels in roots, leaves and shoots, indicating that PreP is required in all plant tissues.

Investigation of the phenotype revealed a difference in growth rate, where the double knockout atprep1 atprep2 grew significantly slower compared to wild type plants. In contrast, the atprep1 and atprep2 mutant grew similarly as the wild type. When we analysed the growth rates of plants grown during 16 hours of light per day we found a statistically significant decrease in the accumulated biomass with about 38 % (dry weight) for the double knockout plants compared to wild type.

A chlorotic phenotype was present in the double knockout and a weaker, but still noticeable chlorotic phenotype in the single knockout atprep1. The chlorotic phenotype was most pronounced during early development in the first true leaves. A pigment comparison study was done and the results showed a clear reduction of the different pigments in double knockout plants compared to wild type plants. The single knockouts did not show any significant reduction. To further analyze the chlorotic phenotype we applied electron microscopy to investigate the ultrastructure of chloroplast and mitochondria in the first true leaves. We found that atprep1 and double knockout chloroplasts contained less starch and poorly developed thylakoids and grana. The PreP2 knockout showed similar structural features as the wild type. The found differences in the chloroplast development could be a result of the observed slower growth rate, which means that the organelle ultrastructure could be in different developmental stages.

Respiratory measurements of mitochondria from the first true leaves showed a significant decrease in respiratory rate for atprep1 atprep2 compared to wild type, measured with malate and glutamate as substrates in the presence of ADP. In addition, the respiratory control ratio (RCR) decreased by 50% indicating partial uncoupling of mitochondria in PreP double knockout plants.

We also investigated the proteolytic activity of the PreP double knockout plants using the F1β(2-54) presequence. Mitochondria and chloroplasts were purified and then sub-fractionated into membrane and soluble fractions. No proteolytic activity could be detected in the mitochondria fractions or in the chloroplast stroma, whilst in the chloroplast membrane fractions a metal dependent activity was detected both in the wild type as well as in the double knockout. This study shows that PreP is important for normal plant growth, but not essential for plant survival in a laboratory environment.

27

Paper II - Binding of divalent cations is essential for the activity of the organellar peptidasome in Arabidopsis thaliana, AtPreP

To investigate the role of the two magnesium sites that were found in the structure of AtPreP1 [108], the six amino acids that coordinate the two magnesium ions were substituted by neutral amino acids. The aspartates were changed to asparagines and threonine was changed to valine. In this way the magnesium binding capacity was depleted and made it possible to compare the activity of the protease with and without bound magnesium at these sites. One of the sites is positioned in the close proximity of the active site and thereby called the inner site or ∆1. In this site the magnesium is coordinated by three aspartate residues, where one (D705) is located in domain three and the other two (D859, D861) in domain four of AtPreP. The other magnesium binding site (composed by D828, T947, and D952) is positioned on the outside surface of the protease in domain four and hence called the outer site or ∆2.

Sequence similarity searches of AtPreP using 19 different sequences showed that the inner magnesium binding site was conserved in plant PrePs. The outer magnesium binding site was only fully conserved in A. thaliana and Physcomitrella patens (moss).

We investigated the metal depleted variants lacking the inner, outer or both metal binding sites, using a native substrate, the mitochondrial presequence from the F1β subunit of ATP synthase, pF1β2-54 at increasing concentrations of Mg2+, Ca2+ and Zn2+ ions. Our results showed that without metal ions no 2+ degradation of pF1β(2-54) could be observed for wild type or any of the variants. The addition of Mg and Ca2+ ions clearly stimulated the catalytic activity of both wild type AtPreP1 and the ∆2 variant at metal concentrations of 1 mM or higher. In contrast, the ∆1 and ∆1,2 variants were only partially active at increasing concentrations of Mg2+ and did not show any activity even at 10 mM of Ca2+. These results suggest that the inner magnesium binding site plays a role for the proteolytic function of 2+ the protease. Addition of Zn did not stimulate the degradation of the pF1β(2-54) peptide. In contrast to Mg2+ and Ca2+, addition of Zn2+ completely inhibited the activity of AtPreP1 already at 1 mM Zn2+.

The proteolytic activity was also tested against 8 other peptides, 9-33 amino acids in length, and an overall degradation pattern could be observed with a higher proteolytic activity in the presence of magnesium ions for all AtPreP1 variants (figure 4-6 in paper II).

The reason for the increased activity when adding MgCl2 is not known, but we speculate that since PreP is very acidic and the enzyme halves are both negatively charged, Mg2+ may assist to stabilize the protein by neutralizing this charge and enable the two halves to come together to form the active site and to enable proteolysis.

28

Paper III - Biochemical studies of SNPs of the mitochondrial Aβ-degrading protease, hPreP

The aim of this project was to study four SNPs located at interesting sites in the hPreP structure, and compare the proteolytic activity with and without Mg2+. The SNPs that were investigated in this study were L116V, F140S, A525D and I924M. SNP L116V is located close to the active site inside the proteolytic chamber and could be in direct contact with the substrate side chains. The substitution to valine may affect substrate binding. SNPs F140S and A525D are found close to the hinge region of PreP and substitutions in this region may affect the opening and closing of the enzyme. Finally, the SNP I924M is located close to the surface of domain 4 about 20 Å from the hinge region and may play a role for protein stability.

We selected four peptides of different lengths and properties to investigate the proteolytic activity of the selected SNP variants: Aβ(1-40), Aβ(1-42), the presequence of ATP synthase F1β subunit (pF1β) and a fluorescent peptide, Substrate V.

Both wild type and hPreP-SNP variants were able to degrade the Aβ peptides. However, we found that the SNP-variants displayed lower activity compared to wild type hPreP. The SNP-variant A525D showed the lowest activity with a degradation of 11 % for the Aβ(1-40) peptide and only 4 % degradation for the Aβ(1-42) peptide compared to wild type hPreP. The activity of the SNP-variants increased in the presence of MgCl2 and reached the same activity as wild type for Aβ(1-40), and almost similar to wild type with Aβ(1-42).

2+ Time titration of the pF1β(2-54) peptide without Mg showed lower activity for the SNP F140S and A525D (about 30% activity compared to wild type hPreP). Interestingly, the activity was fully restored upon addition of 10 mM of Mg2+.

We also studied the kinetics of degradation for the SNP variants compared to wild type hPreP using fluorometry. As substrate we used the fluorogenic Substrate V, which is derived from bradykinin. Substrate V is a nine amino acid long peptide and is positively charged [MCA–Arg-Pro-Pro-Gly-Phe- Ser-Ala-Phe-Lys-(Dnp)–OH, where MCA is (7-methoxycoumarin-4-yl) acetyl and Dnp is 2,4- dinitrophenol]. The peptide harbours a fluorescent MCA group that is efficiently quenched by resonance energy transfer to Dnp. When the peptide is cleaved the two groups can be separated and the quenching is lost, which leads to an increase in fluorescence. Both wild type and hPreP-SNP variants were able to degrade Substrate V, both with and without MgCl2. In contrast, the hPreP(A525D) variant only had 20 % of normal activity in the absence of MgCl2, but interestingly the activity increased to 60 % after the addition of MgCl2.

In summary, these results show that hPreP-SNP variants show reduced activity compared to wild type, and that Mg2+ ions stimulate the proteolytic activity of both wild type hPreP and hPreP-SNP

29 variants. The reason for the increased activity when adding MgCl2 is not known. Similar to AtPreP, hPreP is very acidic and the enzyme halves are both negatively charged. This leads to the hypothesis that Mg2+ may assist to stabilize the protein by neutralizing this charge and enable the two halves to come together to form the active site and enable proteolysis.

30

Future perspectives

Structure and catalytic function of AtPreP1

In future, it would be interesting to solve the structure of wild type AtPreP1 and the inactive variant

AtPreP1 (E80Q) with bound substrate, the F1β presequence (pF1β) or other shorter peptide. In addition to this we would like to solve the structure of the AtPreP1 variant ∆1,2 lacking the two metal binding sites (described above). These structures together with biochemical catalytic activity assays will help to get a better understanding for the catalytic mechanism and the role of non-catalytic metal ions. The initial screening and optimization experiments have led to production of crystals from both wild type

AtPreP1 as for AtPreP1(E80Q)+ F1β presequence, but so far we have not been able to get crystals that diffract better than about 4 Å.

31

Acknowledgements

I would like to thank my supervisor Elzbieta for giving me the opportunity to try the life as a scientist. It has for sure been an interesting time and a journey between despair and joy. From day one you made me feel welcome in the group and the great personal chemistry in the group made me decide to stay. Thank you for your support during these years and for all interesting discussions about science or anything in life.

Stefan, working with you was a huge inspiration and a lot of fun! I was learning so much every day during the DKO project. Shashi, thanks for all the help during my first time in the lab. Nyosha and Anna-Karin, you always kept the EG-lab spirit on top, when you were around anything could and would happen. Catarina, it was fun to work in the dark with you, even if we didn’t see that much we actually got some results. Erika, the Aurora Borealis of the lab, thanks for all the nice moments in the lab and during our travel adventures. Beata and Pedro, thanks for helping out in the lab and for interesting discussions.

Amin, thank you for introducing me to the world of crystallography! It was an interesting but frustrating time. You were my teacher some ten years ago, and to work with you again was a great opportunity for serious discussions about most things. Thank you for always supporting me and believing in me!

Stefan Nordlund, thanks for accepting me as a PhD student at the department. Inger, thanks for teaching me how to teach, and for having answers to my questions.

Thanks to all former EG-lab members that I had the joy to work with, and all other people at the department that I met during my eight years at DBB.

32

References

1. McBride, H.M., M. Neuspiel, and S. Wasiak, Mitochondria: more than just a powerhouse. Curr Biol, 2006. 16(14): p. R551-60. 2. Kennedy, E.P. and A.L. Lehninger, The products of oxidation of fatty acids by isolated rat liver mitochondria. J Biol Chem, 1950. 185(1): p. 275-85. 3. Margulis, L., Origin of Eukaryotic Cells: Evidence and Research Implications for a Theory of the Origin and Evolution of Microbial, Plant, and Animal Cells on the Precambrian Earth. Yale University Press, New Haven, 1970. 4. Searcy, D.G., Metabolic integration during the evolutionary origin of mitochondria. Cell Res, 2003. 13(4): p. 229-38. 5. Lang, B.F., M.W. Gray, and G. Burger, Mitochondrial genome evolution and the origin of eukaryotes. Annu Rev Genet, 1999. 33: p. 351-97. 6. Andersson, S.G., et al., On the origin of mitochondria: a genomics perspective. Philos Trans R Soc Lond B Biol Sci, 2003. 358(1429): p. 165-77; discussion 177-9. 7. Lopez-Juez, E. and K.A. Pyke, Plastids unleashed: their development and their integration in plant development. Int J Dev Biol, 2005. 49(5-6): p. 557-77. 8. Kleine, T., U.G. Maier, and D. Leister, DNA transfer from organelles to the nucleus: the idiosyncratic genetics of endosymbiosis. Annu Rev Plant Biol, 2009. 60: p. 115-38. 9. Berg, O.G. and C.G. Kurland, Why mitochondrial genes are most often found in nuclei. Mol Biol Evol, 2000. 17(6): p. 951-61. 10. Gray, M.W., G. Burger, and B.F. Lang, The origin and early evolution of mitochondria. Genome Biol, 2001. 2(6): p. 1018. 11. Boore, J.L., Animal mitochondrial genomes. Nucleic Acids Res, 1999. 27(8): p. 1767-80. 12. Adams, K.L. and J.D. Palmer, Evolution of mitochondrial gene content: gene loss and transfer to the nucleus. Mol Phylogenet Evol, 2003. 29(3): p. 380-95. 13. Mitchell, P., A chemiosmotic molecular mechanism for proton-translocating adenosine triphosphatases. FEBS Lett, 1974. 43(2): p. 189-94. 14. Timmis, J.N., et al., Endosymbiotic gene transfer: organelle genomes forge eukaryotic chromosomes. Nat Rev Genet, 2004. 5(2): p. 123-35. 15. Green, B.R., Chloroplast genomes of photosynthetic eukaryotes. Plant J, 2011. 66(1): p. 34-44. 16. Krauss, N., Mechanisms for photosystems I and II. Curr Opin Chem Biol, 2003. 7(5): p. 540-50. 17. Schmidt, O., N. Pfanner, and C. Meisinger, Mitochondrial protein import: from proteomics to functional mechanisms. Nat Rev Mol Cell Biol, 2010. 11(9): p. 655-67. 18. Vogtle, F.N., et al., Global analysis of the mitochondrial N-proteome identifies a processing peptidase critical for protein stability. Cell, 2009. 139(2): p. 428-39. 19. Zhang, X.P. and E. Glaser, Interaction of plant mitochondrial and chloroplast signal peptides with the Hsp70 molecular chaperone. Trends Plant Sci, 2002. 7(1): p. 14-21. 20. Pfanner, N. and A. Geissler, Versatility of the mitochondrial protein import machinery. Nat Rev Mol Cell Biol, 2001. 2(5): p. 339-49. 21. Bruch, M.D. and D.W. Hoyt, Conformational analysis of a mitochondrial presequence derived from the F1-ATPase beta-subunit by CD and NMR spectroscopy. Biochim Biophys Acta, 1992. 1159(1): p. 81-93. 22. Chupin, V., et al., Cardiolipin modulates the secondary structure of the presequence peptide of cytochrome oxidase subunit IV: a 2D 1H-NMR study. FEBS Lett, 1995. 373(3): p. 239-44. 23. De Jongh, H.H., The helix nucleation site and propensity of the synthetic mitochondrial presequence of ornithine carbamoyltransferase. Eur J Biochem, 2000. 267(18): p. 5796-804. 24. Karslake, C., et al., 2D NMR and structural model for a mitochondrial signal peptide bound to a micelle. Biochemistry, 1990. 29(42): p. 9872-8.

33

25. MacLachlan, L.K., et al., A spectroscopic study of the mitochondrial transit peptide of rat malate dehydrogenase. Biochem J, 1994. 303 ( Pt 2): p. 657-62. 26. Pfanner, N. and W. Neupert, The mitochondrial protein import apparatus. Annu Rev Biochem, 1990. 59: p. 331-53. 27. Neupert, W. and J.M. Herrmann, Translocation of proteins into mitochondria. Annu Rev Biochem, 2007. 76: p. 723-49. 28. Abe, Y., et al., Structural basis of presequence recognition by the mitochondrial protein import receptor Tom20. Cell, 2000. 100(5): p. 551-60. 29. Brix, J., K. Dietmeier, and N. Pfanner, Differential recognition of preproteins by the purified cytosolic domains of the mitochondrial import receptors Tom20, Tom22, and Tom70. J Biol Chem, 1997. 272(33): p. 20730-5. 30. Lister, R., et al., Functional definition of outer membrane proteins involved in preprotein import into mitochondria. Plant Cell, 2007. 19(11): p. 3739-59. 31. Yano, M., K. Terada, and M. Mori, AIP is a mitochondrial import mediator that binds to both import receptor Tom20 and preproteins. J Cell Biol, 2003. 163(1): p. 45-56. 32. Baker, K.P., et al., A yeast mitochondrial outer membrane protein essential for protein import and cell viability. Nature, 1990. 348(6302): p. 605-9. 33. Lister, R., et al., A transcriptomic and proteomic characterization of the Arabidopsis mitochondrial protein import apparatus and its response to mitochondrial dysfunction. Plant Physiol, 2004. 134(2): p. 777-89. 34. Chacinska, A., et al., Importing mitochondrial proteins: machineries and mechanisms. Cell, 2009. 138(4): p. 628-44. 35. Lister, R., et al., Zinc-dependent intermembrane space proteins stimulate import of carrier proteins into plant mitochondria. Plant J, 2002. 30(5): p. 555-66. 36. Bien, M., et al., Mitochondrial disulfide bond formation is driven by intersubunit electron transfer in Erv1 and proofread by glutathione. Mol Cell, 2010. 37(4): p. 516-28. 37. Kovermann, P., et al., Tim22, the essential core of the mitochondrial protein insertion complex, forms a voltage-activated and signal-gated channel. Mol Cell, 2002. 9(2): p. 363-73. 38. Bonnefoy, N., et al., OXA1, a Saccharomyces cerevisiae nuclear gene whose sequence is conserved from prokaryotes to eukaryotes controls cytochrome oxidase biogenesis. J Mol Biol, 1994. 239(2): p. 201-12. 39. Carrie, C., et al., Conserved and novel functions for Arabidopsis thaliana MIA40 in assembly of proteins in mitochondria and peroxisomes. J Biol Chem, 2010. 285(46): p. 36138-48. 40. Murcha, M.W., et al., The plant mitochondrial protein import apparatus - The differences make it interesting. Biochim Biophys Acta, 2013. 41. Chacinska, A., et al., Distinct forms of mitochondrial TOM-TIM supercomplexes define signal- dependent states of preprotein sorting. Mol Cell Biol, 2010. 30(1): p. 307-18. 42. Bolender, N., et al., Multiple pathways for sorting mitochondrial precursor proteins. EMBO Rep, 2008. 9(1): p. 42-9. 43. Wagner, K., D.U. Mick, and P. Rehling, Protein transport machineries for precursor translocation across the inner mitochondrial membrane. Biochim Biophys Acta, 2009. 1793(1): p. 52-9. 44. Frazier, A.E., et al., Pam16 has an essential role in the mitochondrial protein import motor. Nat Struct Mol Biol, 2004. 11(3): p. 226-33. 45. Truscott, K.N., K. Brandner, and N. Pfanner, Mechanisms of protein import into mitochondria. Curr Biol, 2003. 13(8): p. R326-37. 46. Gakh, O., P. Cavadini, and G. Isaya, Mitochondrial processing peptidases. Biochim Biophys Acta, 2002. 1592(1): p. 63-77. 47. Mossmann, D., C. Meisinger, and F.N. Vogtle, Processing of mitochondrial presequences. Biochim Biophys Acta, 2012. 1819(9-10): p. 1098-106. 48. Taylor, A.B., et al., Crystal structures of mitochondrial processing peptidase reveal the mode for specific cleavage of import signal sequences. Structure, 2001. 9(7): p. 615-25.

34

49. Rawlings, N.D., D.P. Tolle, and A.J. Barrett, MEROPS: the peptidase database. Nucleic Acids Res, 2004. 32(Database issue): p. D160-4. 50. Kitada, S., et al., A putative metal-binding site in the beta subunit of rat mitochondrial processing peptidase is essential for its catalytic activity. J Biochem, 1995. 117(6): p. 1148-50. 51. Yang, M., et al., Import of proteins into yeast mitochondria: the purified matrix processing protease contains two subunits which are encoded by the nuclear MAS1 and MAS2 genes. EMBO J, 1988. 7(12): p. 3857-62. 52. Braun, H.P., et al., The general mitochondrial processing peptidase from potato is an integral part of cytochrome c reductase of the respiratory chain. Embo J, 1992. 11(9): p. 3219-27. 53. Braun, H.P. and U.K. Schmitz, The bifunctional cytochrome c reductase/processing peptidase complex from plant mitochondria. J Bioenerg Biomembr, 1995. 27(4): p. 423-36. 54. Eriksson, A.C., S. Sjoling, and E. Glaser, Characterization of the bifunctional mitochondrial processing peptidase (MPP)/bc1 complex in Spinacia oleracea. J Bioenerg Biomembr, 1996. 28(3): p. 285-92. 55. Glaser, E. and P. Dessi, Integration of the mitochondrial-processing peptidase into the cytochrome bc1 complex in plants. J Bioenerg Biomembr, 1999. 31(3): p. 259-74. 56. Nunnari, J., T.D. Fox, and P. Walter, A mitochondrial protease with two catalytic subunits of nonoverlapping specificities. Science, 1993. 262(5142): p. 1997-2004. 57. Gavel, Y. and G. von Heijne, Cleavage-site motifs in mitochondrial targeting peptides. Protein Eng, 1990. 4(1): p. 33-7. 58. Isaya, G., et al., Cleavage of precursors by the mitochondrial processing peptidase requires a compatible mature protein or an intermediate octapeptide. J Cell Biol, 1991. 113(1): p. 65-76. 59. Naamati, A., et al., Dual targeting of Nfs1 and discovery of its novel processing enzyme, Icp55. J Biol Chem, 2009. 284(44): p. 30200-8. 60. Jarvis, P., Targeting of nucleus-encoded proteins to chloroplasts in plants. New Phytol, 2008. 179(2): p. 257-85. 61. Flores-Perez, U. and P. Jarvis, Molecular chaperone involvement in chloroplast protein import. Biochim Biophys Acta, 2013. 1833(2): p. 332-40. 62. May, T. and J. Soll, 14-3-3 proteins form a guidance complex with chloroplast precursor proteins in plants. Plant Cell, 2000. 12(1): p. 53-64. 63. Schnell, D.J., F. Kessler, and G. Blobel, Isolation of components of the chloroplast protein import machinery. Science, 1994. 266(5187): p. 1007-12. 64. Tranel, P.J., et al., A component of the chloroplastic protein import apparatus is targeted to the outer envelope membrane via a novel pathway. EMBO J, 1995. 14(11): p. 2436-46. 65. Hinnah, S.C., et al., The chloroplast protein import channel Toc75: pore properties and interaction with transit peptides. Biophys J, 2002. 83(2): p. 899-911. 66. Tsai, J.Y., et al., Structural characterizations of the chloroplast translocon protein Tic110. Plant J, 2013. 75(5): p. 847-57. 67. Trosch, R. and P. Jarvis, The stromal processing peptidase of chloroplasts is essential in Arabidopsis, with knockout mutations causing embryo arrest after the 16-cell stage. PLoS One, 2011. 6(8): p. e23039. 68. Richter, S. and G.K. Lamppa, Stromal processing peptidase binds transit peptides and initiates their ATP-dependent turnover in chloroplasts. J Cell Biol, 1999. 147(1): p. 33-44. 69. Carrie, C., E. Giraud, and J. Whelan, Protein transport in organelles: Dual targeting of proteins to mitochondria and chloroplasts. FEBS J, 2009. 276(5): p. 1187-95. 70. Peeters, N. and I. Small, Dual targeting to mitochondria and chloroplasts. Biochim Biophys Acta, 2001. 1541(1-2): p. 54-63. 71. Carrie, C. and I. Small, A reevaluation of dual-targeting of proteins to mitochondria and chloroplasts. Biochim Biophys Acta, 2013. 1833(2): p. 253-9. 72. Garcia-Lorenzo, M., et al., Protease gene families in Populus and Arabidopsis. BMC Plant Biol, 2006. 6: p. 30.

35

73. Sakamoto, W., Protein degradation machineries in plastids. Annu Rev Plant Biol, 2006. 57: p. 599-621. 74. Alexopoulos, J.A., A. Guarne, and J. Ortega, ClpP: a structurally dynamic protease regulated by AAA+ proteins. J Struct Biol, 2012. 179(2): p. 202-10. 75. Peltier, J.B., et al., Clp protease complexes from photosynthetic and non-photosynthetic plastids and mitochondria of plants, their predicted three-dimensional structures, and functional implications. J Biol Chem, 2004. 279(6): p. 4768-81. 76. Adam, Z., A. Rudella, and K.J. van Wijk, Recent advances in the study of Clp, FtsH and other proteases located in chloroplasts. Curr Opin Plant Biol, 2006. 9(3): p. 234-40. 77. Ito, K. and Y. Akiyama, Cellular functions, mechanism of action, and regulation of FtsH protease. Annu Rev Microbiol, 2005. 59: p. 211-31. 78. Santos, D. and D.F. De Almeida, Isolation and characterization of a new temperature-sensitive cell division mutant of Escherichia coli K-12. J Bacteriol, 1975. 124(3): p. 1502-7. 79. Wagner, R., H. Aigner, and C. Funk, FtsH proteases located in the plant chloroplast. Physiol Plant, 2012. 145(1): p. 203-14. 80. Sakamoto, W., et al., Coordinated regulation and complex formation of yellow variegated1 and yellow variegated2, chloroplastic FtsH metalloproteases involved in the repair cycle of photosystem II in Arabidopsis thylakoid membranes. Plant Cell, 2003. 15(12): p. 2843-55. 81. Sokolenko, A., et al., The gene complement for proteolysis in the cyanobacterium Synechocystis sp. PCC 6803 and Arabidopsis thaliana chloroplasts. Curr Genet, 2002. 41(5): p. 291-310. 82. Janska, H., M. Kwasniak, and J. Szczepanowska, Protein quality control in organelles - AAA/FtsH story. Biochim Biophys Acta, 2013. 1833(2): p. 381-7. 83. Besche, H. and P. Zwickl, The Thermoplasma acidophilum Lon protease has a Ser-Lys dyad active site. Eur J Biochem, 2004. 271(22): p. 4361-5. 84. Botos, I., et al., The catalytic domain of Escherichia coli Lon protease has a unique fold and a Ser-Lys dyad in the active site. J Biol Chem, 2004. 279(9): p. 8140-8. 85. Rotanova, T.V., et al., Slicing a protease: structural features of the ATP-dependent Lon proteases gleaned from investigations of isolated domains. Protein Sci, 2006. 15(8): p. 1815- 28. 86. Adam, Z., et al., Chloroplast and mitochondrial proteases in Arabidopsis. A proposed nomenclature. Plant Physiol, 2001. 125(4): p. 1912-8. 87. Kikuchi, M., et al., Proteomic analysis of rat liver peroxisome: presence of peroxisome-specific isozyme of Lon protease. J Biol Chem, 2004. 279(1): p. 421-8. 88. Ostersetzer, O., et al., Multiple intracellular locations of Lon protease in Arabidopsis: evidence for the localization of AtLon4 to chloroplasts. Plant Cell Physiol, 2007. 48(6): p. 881-5. 89. Urban, S. and S.W. Dickey, The rhomboid protease family: a decade of progress on function and mechanism. Genome Biol, 2011. 12(10): p. 231. 90. Mayer, U. and C. Nusslein-Volhard, A group of genes required for pattern formation in the ventral ectoderm of the Drosophila embryo. Genes Dev, 1988. 2(11): p. 1496-511. 91. Bier, E., L.Y. Jan, and Y.N. Jan, rhomboid, a gene required for dorsoventral axis establishment and peripheral nervous system development in Drosophila melanogaster. Genes Dev, 1990. 4(2): p. 190-203. 92. Wasserman, J.D., S. Urban, and M. Freeman, A family of rhomboid-like genes: Drosophila rhomboid-1 and roughoid/rhomboid-3 cooperate to activate EGF receptor signaling. Genes Dev, 2000. 14(13): p. 1651-63. 93. Kanaoka, M.M., et al., An Arabidopsis Rhomboid homolog is an intramembrane protease in plants. FEBS Lett, 2005. 579(25): p. 5723-8. 94. Schuhmann, H. and I. Adamska, Deg proteases and their role in protein quality control and processing in different subcellular compartments of the plant cell. Physiol Plant, 2012. 145(1): p. 224-34.

36

95. Lipinska, B., et al., Identification, characterization, and mapping of the Escherichia coli htrA gene, whose product is essential for bacterial growth only at elevated temperatures. J Bacteriol, 1989. 171(3): p. 1574-84. 96. Strauch, K.L., K. Johnson, and J. Beckwith, Characterization of degP, a gene required for proteolysis in the cell envelope and essential for growth of Escherichia coli at high temperature. J Bacteriol, 1989. 171(5): p. 2689-96. 97. Clausen, T., C. Southan, and M. Ehrmann, The HtrA family of proteases: implications for protein composition and cell fate. Mol Cell, 2002. 10(3): p. 443-55. 98. Lipinska, B., M. Zylicz, and C. Georgopoulos, The HtrA (DegP) protein, essential for Escherichia coli survival at high temperatures, is an endopeptidase. J Bacteriol, 1990. 172(4): p. 1791-7. 99. Kley, J., et al., Structural adaptation of the plant protease Deg1 to repair photosystem II during light exposure. Nat Struct Mol Biol, 2011. 18(6): p. 728-31. 100. Krojer, T., et al., Structural basis for the regulated protease and chaperone function of DegP. Nature, 2008. 453(7197): p. 885-90. 101. Kapri-Pardes, E., L. Naveh, and Z. Adam, The thylakoid lumen protease Deg1 is involved in the repair of photosystem II from photoinhibition in Arabidopsis. Plant Cell, 2007. 19(3): p. 1039- 47. 102. Kmiec, B., et al., Organellar oligopeptidase (OOP) provides a complementary pathway for targeting peptide degradation in mitochondria and chloroplasts. Proc Natl Acad Sci U S A, 2013. 110(40): p. E3761-9. 103. Keech, O., P. Dizengremel, and P. Gardestrom, Preparation of leaf mitochondria from Arabidopsis thaliana. Phys Plant, 2005. 124: p. 403-409. 104. Weigel, D. and J. Glazebrook, Arabidopsis: A Laboratory Manual. Cold Spring Harbor Laboratory Press, 2002: p. 217-219. 105. Becker, A.B. and R.A. Roth, An unusual active site identified in a family of zinc . Proc Natl Acad Sci U S A, 1992. 89(9): p. 3835-9. 106. Stahl, A., et al., Two novel targeting peptide degrading proteases, PrePs, in mitochondria and chloroplasts, so similar and still different. J Mol Biol, 2005. 349(4): p. 847-60. 107. Alikhani, N., et al., Targeting capacity and conservation of PreP homologues localization in mitochondria of different species. J Mol Biol, 2011. 410(3): p. 400-10. 108. Johnson, K.A., et al., The closed structure of presequence protease PreP forms a unique 10,000 Angstroms3 chamber for proteolysis. Embo J, 2006. 25(9): p. 1977-86. 109. Moberg, P., et al., NMR solution structure of the mitochondrial F1beta presequence from Nicotiana plumbaginifolia. J Mol Biol, 2004. 336(5): p. 1129-40. 110. Bruce, B.D., Chloroplast transit peptides: structure, function and evolution. Trends Cell Biol, 2000. 10(10): p. 440-7. 111. Aleshin, A.E., et al., Crystal and solution structures of a prokaryotic M16B peptidase: an open and shut case. Structure, 2009. 17(11): p. 1465-75. 112. Nicolay, K., F.D. Laterveer, and W.L. van Heerde, Effects of amphipathic peptides, including presequences, on the functional integrity of rat liver mitochondrial membranes. J Bioenerg Biomembr, 1994. 26(3): p. 327-34. 113. Falkevall, A., et al., Degradation of the amyloid beta-protein by the novel mitochondrial peptidasome, PreP. J Biol Chem, 2006. 281(39): p. 29096-104. 114. Anandatheerthavarada, H.K. and L. Devi, Amyloid precursor protein and mitochondrial dysfunction in Alzheimer's disease. Neuroscientist, 2007. 13(6): p. 626-38.

37