2. .. 1 :25... 27. .1.) (.132 c .. .. 3.9.va . 1. _ . 3&2»??? $22 1?...» . .2 . E. . r2 5...... t 1 . “N. u. .2 y a. 2.2.3. .zv5zs25h. .39.... l 2 q 5 121.15.... . 12$ .1... 3:551 . I. .3 n ‘1 . ‘IA ul¢n ‘55- ....3 .1: .7. 1 12-1.1.2... .l .‘l? & 1 1...» .2... .2. .25.. 1.21... .- s 2 .5: a)! 2...... 2:13.. 1 5r. :2. vav»...§.~.€$ 050139.11.- l“ ‘?flv".filmht. .. It "41. . .22 a . 11):] 3...... .. unannw... .12., 11.22.41.113? unHqu....§: fiafi? I O . A1... 1. .3 . .933 I... .21..- .22. :i 1‘ .41;- 1") ‘3} , 2: 7. 2.! 13.2.1 12...: .1... .1. 2... Irinfib h6!.1.rll$k$ . ,‘lb‘r urns}... ”.12.... .2 . .2,» 3.2.11: 5.3%..5 5.7? in. u 2 35.2.2223: , 43“ $2 . .21., 2.0.1.5310; . r n 5 , 1'5“..- 75L. nuitifikgmv . 1 1 Enltévznu)V SI! 13.2.2.1? . 1 . .15.. 22.1.1.3 ”1.2.1. 1222.}. .5.» 1.. 2 {2.1.5.2. 4., an km»... 23.5.. «3 , . 41.11.... . 3111!: . 22 ark}; 2 a 1, 45- 4 1 k 5 2 3.3.54.3 . .. i e. 1 more ‘ LIBRARY 2% Michigan State University This is to certify that the thesis entitled IDENTIFICATION AND CHARACTERIZATION OF TRYPANOSOMA BRUCEI PPR PROTEINS, PUTATIVE MITOCHONDRIAL RNA METABOLISM PROTEINS presented by Melissa Kay Mingler has been accepted towards fulfillment of the requirements for the Master of degree in WSW and Molecular Science A / Biology : 7 -.-.-.-¢--n-o-n-I-o-o--I-o-n-o---1-.-- MSU is an Affinnative Action/Equal Opportunity Institution 4. __ 1+ “ '_—"—' * ‘. PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECAU£D with earlier due date if requested. DATE DUE DATE DUE DATE DUE 2/os'c/c'i—RC/oa't'e—oueindd-"—p. 15 ‘ IDENTIFICATION AND CHARACTERIZATION OF TRYPANOSOMA BRUCEI PPR PROTEINS, PUTATIVE MITOCHONDRIAL RNA METABOLISM PROTEINS BY Melissa Kay Mingler A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTERS OF SCIENCE Department of Biochemistry and Molecular Biology 2005 ABSTRACT IDENTIFICATION AND CHARACTERIZATION OF TRYPANOSOMA BRUCEI PPR PROTEINS, PUTATIVE MITOCHONDRIAL RNA METABOLISM PROTEINS BY Melissa Kay Mingler A new class of proteins, characterized by Pentatricopeptide Repeat (PPR) motifs, have been identified recently in plants. These proteins contain multiple 35-amino acid repeats that are proposed to form a superhelix capable of binding a strand of RNA. All PPR proteins characterized to date appear to be involved in RNA metabolism in organelles. Comparative genomic studies show that while there are 23 PPR proteins within Trypanosoma brucez; plants contain over 450 PPR proteins. In contrast, eukaryotes only contain 2-6 PPR proteins. Studies began with bioinformatics that characterized a 7'. brucei-speciflc PPR motif. One of the putative 72 brucei mitochondrial PPR proteins, TbPPR1, was further characterized using RNA interference (RNAi). TbPPRI is predicted to be mitochondrially targeted and contains 14 predicted PPR motifs, with the majority occurring in tandem. RNAi experiments designed to knockdown expression of TbPPRl show a repeatable slow growth phenotype. The mitochondrial mRNA levels from the RNAi experiments were studied using Northern blot analysis and with a “poison primer“ extension assay, and several messages showed an altered expression pattern. Through these studies we hope to identify relevant RNA processing factors of mitochondrial messages. ACKNOWLEDGMENTS Thank you to Donna Koslowsky, Ph.D. who supported and instructed me through my graduate studies. Also a thank you to Sandra L. Clement, Ph.D. for instructing me in many of the lab techniques used in this thesis as well as for all her helpful advice. I would like to acknowledge Kaillathe Padmanabhan, Ph.D. for his help in using the HMMER 2.2 package for my bioinformatics studies and all his other computer expertise. I would also like to acknowledge Annette Thelen, Ph.D. for all her helpful advise on the Real Time PCR studies and Shirley Owens for her help in the fluorescent microscopy studies. iii TABLE OF CONTENTS LIST OF TABLES ...................................................................................................... V LIST OF FIGURES ................................................................................................... VI CHAPTER 1: INTRODUCTION ................................................................................. 1 Introduction .................................................................................................. 2 Trypanosomes: Lifecycle, and Medical Significance ................................... 2 Trypanosoma brucei Metabolism ................................................................. 6 Mitochondrial Genome and mRNA Editing ................................................... 8 Mitochondrial Regulation ............................................................................ 15 PPR Proteins ................................................................................................ 20 Overview ..................................................................................................... 26 CHAPTER 2: BIOINFORMATICS ............................................................................ 28 Introduction ................................................................................................ 29 Materials and Methods ................................................................................ 30 Results .......................................................................................................... 31 PPR Consensus Motifs Identified with HMMER and MEME .............. 31 Protein Characterization ................................................................... 34 Discussion ................................................................................................... 42 CHAPTER 3: RNAi DATA ....................................................................................... 50 Introduction ................................................................................................. 51 Materials and Methods ................................................................................ 52 RNAi constructs ................................................................................ 52 Transfection, RNAi induction, and Growth Curve ............................ 56 Northern Analyses ............................................................................ 57 Poison Primer Extension ................................................................... 57 Results ......................................................................................................... 58 RNAi .................................................................................................. 58 Evaluation of TbPPR1 Depletion effects on Mitochondrial RNA Transcripts ............................................................................. 65 Poison Primer Extension ................................................................... 72 Discussion and Conclusion .......................................................................... 76 Future Work ..................................................................................... 78 BIBLIOGRAPHY ...................................................................................................... 80 iv LIST OF TABLES . Gene overlap of 72 brucei Maxicircle .......................................................... 14 . Regulation of 72 brucei Mitochondrial Expression ..................................... 19 . Eukaryotic PPR proteins .............................................................................. 23 . PPR proteins are specific to Eukaryotes ..................................................... 25 . 72 brucei PPR proteins ................................................................................. 39 . Orthologs of 72 brucei PPR proteins ........................................................... 41 LIST OF FIGURES 1. Trypanosoma brucei life cycle ...................................................................... 5 2. Linearized maxicircle map ........................................................................... 11 3. 7i brucei minicircle ...................................................................................... 12 4. Position of the maxicircle encoded gMURFII-Z gene in 7? brucei. ............. 13 5. 72 bmcei specific PPR motif ........................................................................ 33 6. TbPPR1 PPR motif sequences ..................................................................... 36 7. TbPPR1 putative RNA contacting residues ................................................. 44 8. Family of a—helical repeat proteins ............................................................. 47 9. TbPPR1 PPR domains .................................................................................. 53 10.pZ]M, T. bruceidsRNA expression vector .................................................. 55 11.TbPPR1 RNAi Growth Curves ...................................................................... 61 12. Control for Tetracycline Exposure ............................................................... 62 13. Northern blot of TbPPR1 RNAi Trial A: TbPPR1 and dsRNA expression....64 14. Northern blot of TbPPR1 RNAi Trial C/D: 125 and 9s rRNA expression ..... 66 15.Northern blot of TbPPR1 RNAi Trial C/D: Complex I expression ............... 69 16. Northern blot of TbPPR1 RNAi Trial C/D: CYb expression ......................... 71 17. Poison Primer Extension of CYb on TbPPR1 RNAi Total RNA .................... 75 vi CHAPTER 1 INTRODUCTION Introduction Pentatricopeptide Repeat (PPR) proteins are an important family of a- helical repeat proteins involved in several aspects of organellar RNA post- transcriptional processing in many different eukaryotic organisms. A large number of these proteins have been identified in Trypanosoma brucei. We plan to investigate their role in mitochondrial mRNA processing, stability or expression. This chapter introduces the organism 72 brucei its significance for study, and its mitochondrial biogenesis, as well as the family of PPR proteins. Trypanosomes: Lifecycle, Medical and Economical Significance. Trypanosoma brucei are parasitic protozoa that cause African Sleeping Sickness in humans and Nagana in cattle. They threaten over 60 million people in 36 different countries in sub-Saharan Africa. In 1999, 45,000 cases of African Sleeping Sickness were diagnosed. However, the World Health Organization (WHO) estimates that the number of people affected is ten times greater due to the lack of screening. There are several villages in Angola, the Democratic Republic of Congo and Southern Sudan where Sleeping Sickness is the first or second greatest cause of mortality, even ahead of HIV/AIDS (www.who.int). 7'. brucei is not only detrimental to the health of the people of sub-Saharan Africa, but also to their well-being. Infection in animals causes decimation of cattle and abandonment of fertile land to avoid the disease. The symptoms in the early stages of the disease are fever, headaches, pains in the joints, and itching. The second stage is the neurological phase where 7? brucei crosses the blood-brain barrier into the central nervous system. The symptoms are confusion, sensory disturbances, poor coordination, and disturbances in the sleep cycle that all can culminate in fatality. Treatment in the initial stage is more effective and safer. The only second stage drug on the market today is Melarsoprol, which is an arsenic derivative with severe and deadly side effects, killing 4-12% of all patients who receive it [3]. Elfornithine also treats late stage infections, but only those caused by one of the subspecies. Elfornithine’s maker, Adventis, has only guaranteed production of the injectable form until the year 2005 (www.who.int). * The 7'. brucei parasite is transmitted through the bite of the Tsetse fly, Glossina, when it takes a blood meal from a vertebrate host. It then proliferates in waves in the host’s bloodstream where it evades the immune system by continually changing its antigenic coat of variant surface glycoproteins [4-6]. There are two life forms in the bloodstream. The long, slender form is able to undergo replication. The short, stumpy forms are nonreplicative and begin to change in metabolism in preparation to enter the next lifecycle stage in the tsetse fly. When the tsetse fly takes a blood meal from an infected individual, the short and stumpy bloodstream form will differentiate into the procyclic form in the mid-gut of the insect. This is the form that we study in the lab. The parasites then travel up to the salivary glands of the fly and differentiate again into the epimastigotes and then metacyclics, which are then delivered to the mammals when the tsetse takes a bloodmeal (Fig. 1). Trypanosoma brucei lifecycle. tsetse fly mammal procyclic bloodstream form Figure 1. The lifecycle of Trypanosoma brucei. The light gray areas represent the mitochondria that have great morphological changes between lifecycle stages. The large circle with the dot is the nucleus and the smaller oval with the dot is the kinetoplast. Obtained from homepage.mac.com/mfield/ lab/Images/Iifecycle.gif. FIGURE 1 Trypanosoma brucei Metabolism Metabolism in 72 brucei is quite complex due to the digenetic life cycle. The bloodstream form parasite, in the mammalian host, depends on glucose from the hosts’ bloodstream for its energy production and secretes pyruvate [7, 8]. The procyclic or insect form parasite uses proline, glucose and threonine from the insects’ gut for its energy production and secretes succinate, acetate, lactate, alanine, and C02 [9, 10]. The respiration rate in the bloodstream form of Trypanosoma brucei is quite high, 50-fold higher than any eukaryotic cell [11]. Metabolism in the bloodstream form occurs in a unique peroxisome-like organelle called the glycosome. Glycosomes have a single phospholipid bilayer with an electron dense proteinaceous matrix and no DNA. The glycosome contains glycolytic enzymes and enzymes of peroxide metabolism, fatty acid oxidation, and ether lipid biosynthesis. Metabolism occurs here through substrate level phosphorylation of glucose that produces 2 moles of ATP for every mole of glucose consumed. No net ATP or NADH is produced in the glycosome, as the ATP is actually produced in the cytoplasm by pyruvate kinase [10]. In the mitochondria of the bloodstream forms, there is no TCA cycle and most of the cytochromes are not expressed [12]. The intermembrane space of the mitochondria does contain a glycerol-3-phosphate dehydrogenase and a terminal alternative oxidase that are principal in the glycerol-3-phosphate and dihydroxyacetone phosphate shunt between the glycosome and the mitochondria. This shunt maintains the NAD+/NADH balance within the glycosome. There is also evidence of an active Complex I in the mitochondria that may transfer electrons via the alternative oxidase [13]. The Fo/Fl-ATP synthase of the mitochondria is responsible for maintaining a proton gradient across the mitochondrial membrane in the bloodstream form trypanosomes [14]. The metabolism in the procyclic form is quite different. At low levels of glucose, oxidative phosphorylation within the mitochondria is essential. The mitochondria contain and use an incomplete TCA cycle. The succinyl-CoA synthetase of the TCA has been found to be essential as the last step in both the glucose and proline degradation pathways [15-17]. The TCA cycle is thought to feed into the electron transport chain when Complex II transfers the electrons from succinate of the TCA cycle to ubiquinone of the electron transport chain, thereby skipping Complex I in the procyclic forms [10, 18]. The procyclic form contains two terminal oxidases in its electron transport chain, the cytochrome c oxidase and the plant-like alternative oxidase. The cytochrome c oxidase is active in the mitochondrial oxidative phosphorylation, but the role of the alternative oxidase within the procyclic form is unknown. The protein levels of the alternative oxidase are lower in the procyclic then they are in the bloodstream form [19]. There is some evidence that the alternative oxidase is responsible for decreasing the reactive oxygen species (ROS) in procyclics [20]. The Fo/Fl-ATP synthase is the principal site of ATP generation. Metabolism Summary. The glycosome is highly active in the bloodstream form of the parasite producing energy through substrate level phosphorylation. The role of the mitochondria in the bloodstream form is minor in that it contains a glycerol-3-phosphate dehydrogenase and the alternative oxidase that are important in maintaining the NAD+/NADH ratio in the glycosome, has some Complex I activity and uses the Fo/Fl-ATP synthase to maintain the proton gradient. Mitochondrial function in the procyclic is increased with its use of the oxidative phosphorylation, substrate level phosphorylation, and TCA cycle. Mitochondrial Genome and mRNA Editing The mitochondrial genome of 72 brucei consists of ~50 copies of 3 ~23 kb maxicircle (Fig. 2) and 5,000-10,000 1.0 kb minicircles (~300 sequence families)(Fig. 3). These circles are topologically interlocked forming a network in the form of a disk called the kinetoplast DNA (kDNA). The maxicircles are analogous to mtDNA of other organisms. They encode 18 messenger RNAs (mRNAs), two guide RNAs (gRNA), a large and small ribosomal RNA (rRNA), but surprisingly, there are no transfer RNAs (tRNA) encoded in the maxicircles. Though the maxicircles are ~23 kb in size, only about 15 kb of this is actually coding region (Fig. 2). The remaining 8 kb of the maxicircle is called the variable region and it is thought that transcription of the polycistronic message begins somewhere in this region [21-23] (Fig. 2). Michelotti et al. have shown the existence of a transient precursor element, which would place the transcription start site about 1,200 nt upstream of the 125 rRNA mature 5' end [22], but the exact location is still unknown. The only mapped transcription start site on the maxicircle is for gMURFZ-II, one of the two gRNAs that are encoded on the 72 brucei maxicircle. The gMURFZ-II gene is found completely within the 5’ end of the gene ND4, introducing more complexity to the transcription and processing of the maxicircle [24] (Fig. 4). Minicircle transcription is somewhat different; each of the three gRNAs on a minicircle are primary transcription products located within cassettes of imperfect inverted 18-bp repeats [25, 26]. The 5’ ends of many gRNAs have been mapped 29-33 bp from the upstream repeat [26, 27]. The 18-bp inverted repeats have been implicated in playing a role in transcription, but the few gRNA genes located outside of these 18-bp repeats have also been found to be primary transcripts [24, 28, 29]. Even though each gRNA gene has the ability to initiate transcription, it appears that minicircle gRNA genes can be transcribed polycistronically [30], but the majority of the gRNAs found 'n the cell have 5’ di- or tri-phosphates, indicating they are not processed at their 5’ ends [24, 26]. Though the transcription start site of the maxicircle is unknown, it is known that transcription of the maxicircle results in polycistronic transcripts from which the individual RNAs are then processed out. The 5’ and 3’ ends of many maxicircle genes have been mapped. The coding region is very compact, with the majority of the genes overlapping [3 1](T able 1) at the 5’ and 3’ ends. Thus processing of one message will often result in the destruction of its neighbor(s) [31, 32]. The control of which of the neighboring messages is processed out and translated remains unknown. Extensive RNA processing of maxicircle transcripts must occur before a translatable message is obtained, including endonucleolytic cleavage, polyadenylation and RNA editing [22, 24, 31]. Kinetoplastid RNA editing inserts or deletes uridylates into the mRNA transcripts, forming the correct open reading frames, start and stop codons for translation. Once transcribed, many of the mRNAs are edited by gRNAs, primary transcripts encoded mainly by the minicircles [27] (Fig. 3). These gRNAs are present in both life cycle stages, though almost nothing is known about gRNA transcription, processing or transcript level regulation. 10 Linearized Maxicircle Map. Coding region (~15 kb‘r Variable region (~7_4 kb) N07 CYb cna MURF2 N04 93 CO3 A6 CO2 N05 3’ 8. 5’ ends overlapping Figure 2. Linearized maxicircle map (coding region). 125 and 9s = rRNA subunits; ND=NADH dehydrogenase; CO=cytochrome oxidase; CYb=Cytochrome b; MURF=maxicircle unidentified reading frame; CR=C—rich region; RSP12=ribosomal protein 12. Overlapping regions are shown in gray, unedited genes are striped, extensively edited genes are in black, and the dimpled genes are genes that go through some mRNA editing. Courtesy of Sandra Clement. FIGURE 2 11 T. brucei Minicircle BEND ' Ol‘i 4 // Minicircle // 0° / / P. % lkb % 5 s It .. ’ . ' =18bp inverted repeats lefi Figure 3. 73 brucei minicircle. The gray arrows represent the 18bp inverted repeats that flank most gRNA genes. The striped regions are the gRNAs, typically 3 per minicircle. The oval, ori, represents the origin of replication and the box indicates the bend present in the minicircle DNA structure. The black arrows indicate the direction of the gRNA transcription. FIGURE 3 12 Position of the maxicircle encoded gMURFII—Z gene in 11 brucei 5 ’ ND4 5 ’ gMurfII-Z TAAGAAGG—-)AAATTT:>ATAGAAAGCACAAAAATAAAATTAAATTAGAGTAATTGAATGTTAAAATTiAAATT Figure 4. Position of the maxicircle encoded gMurfII-Z gene in T. brucei. The mapped 5’ ends of both ND4(—->) and gMurfII-Z (:>) are indicated with arrows pointing in the direction of transcription. The 5’ end of ND4 is processed, whereas the 5’ end of gMurfII-2 is a primary transcript. The arrow pointing down indicates the position of gRNA mapped 3’ end. The ATG start codon for ND4 is shown in bold. Courtesy of Donna Koslowsky. FIGURE 4 13 Gene overlap of T. brucei Maxicircle. Opposite Strands Orientation Overlap Same Strand Orientation Overlap N DB/N 09 373' 49nts ND7/COI l| 3'/5' 32nts A6/ M U R Fl 3'/3' 57nts COI l I/Cyb 3'/5' 3nts MU R Fl/C R3 5'/5' 61 nts Cyb/A6 3'/5' 0nts ORB/ND1 3'/3' 41 nts COI l/MU RFll 3'/5' 31 nts ND1/COII 5'/5' 129nts COI/CR4 5'/3' 56nts MU R Fl I/COI 3'/3' 76nts RSP12/ND5 3'/5' 39nts C R4/N D4 5'/5' 8nt intergen N D4/C R5 3'/3' 44nts CR5/RSP12 5'/5' 50nts Table 1. Table representing gene overlap on the maxicircle for border regions of different genes. The left half of the table lists neighboring genes on opposite strands with their orientation and overlap region. The right half of the table lists the neighboring genes on the same strand with their orientation and overlap region. Courtesy of Donna Koslowsky. TABLE 1 14 Mitochondrial Regulation During the life cycle, the changes in energy metabolism, discussed previously, are accompanied by distinct changes in mitochondrial gene expression. Steady state levels of mitochondrial mRNAs and rRNAs, polyadenylation of mitochondrial mRNAs, and editing of these same RNAs are also developmentally regulated in a transcript specific manner[33-36]. mRNA Editing 7: brucei mitochondrial mRNA editing is regulated in a transcript specific manner between the two life cycle stages. This regulation is not controlled by the gRNA since they are present in both life cycle stages [27]. The Complex I edited mRNAs NADH Dehydrogenase 8 (ND8) and the 3’ editing domain of ND7 are only fully edited in the bloodstream forms, whereas ND9 edited mRNA is similar in abundance in both life cycle stages [32, 37-39]. Ribosomal protein subunit 12, RP512, has a higher level of edited mRNA in the bloodstream form [32] and C-rich region 4 (CR4) is only fully edited in the bloodstream form, just like ND7 and N08 [40]. In contrast, the edited form of Cytochrome Oxidase III (COIII), Cytochrome b (CYb), and C011 occurs to a higher extent in procyclic than bloodstream forms [36, 39, 41-44]. ATP Synthase 6 (A6) and Maxicircle Unidentified Reading Frame II (MURFII) have similar levels of the edited form 15 between life cycle stages [43, 45]. MURFI, COI, ND4, and ND5 are never edited (Table 2). te e mRNA evels Though the transcription rates of the mitochondrial RNAs are unchanged between the two life cycle forms, the steady state levels of many of the mitochondrial transcripts differ in a transcript specific manner [22]. Steady state levels of the mitochondrial 125 and 95 rRNAs are 30 fold higher in the procyclic form compared to the long, slender bloodstream form [21]. Similarly, transcripts for members of both Complex III (CYb) and Complex IV (C011 and COI) are also upregulated in procyclic forms [29, 46]. In contrast, Complex I subunits NDB-S and ND7-9, MURFI, and RP512 are elevated in the bloodstream forms of 72 brucei [32, 35, 37, 38, 47, 48]. This supports the suggestion that the procyclics bypass Complex I in the electron transport chain [41]. Other transcripts, A6, N01, and MURFII are constitutively expressed and show no difference in transcript levels between the two forms [35, 45, 46] (Table 2). P l en l ti n All protein coding mRNAs in the mitochondria of 72 brucel; with the exception of ND5, occur in 2 size classes that differ by 120-200 nucleotides in length, a difference which cannot be accounted for by RNA editing [33, 34, 36, 49]. RNase H digestion using oligo(dT) showed that this can be accounted for 16 by the addition of two separate poly(A) tail lengths, short (20 nucleotides) and long (120-200 nucleotides)[46]. Evidence is beginning to emerge that this poly(A) tail is involved in mitochondrial mRNA decay in T. brucei. Unedited RP512 is targeted for degradation with the addition of a poly(A) tail, but has a much longer half life if a poly(A) tail is not added. In contrast, an edited RP512 mRNA or even a 10% partially 3’ edited RP512 mRNA is quickly degraded if it does not possess a poly(A) tail. The longer (120-200 nt) poly(A) tails on edited RP512 mRNAs actually are degraded faster then the shorter (20 nt) tails, but slower then a non-poly(A) tail edited mRNA. This suggests a role of both the 6119 edited portion of the mRNA and the poly(A) tail in stabilization [50]. This pathway of mRNA degradation is also dependent on the addition of UTP [51]. This brings up the question of why the larger size transcript (larger poly(A) tail) class consists of mainly edited mRNAs [32, 37, 46, 47, 52, 53] if a longer poly(A) tail is more destabilizing? Complex I mRNA ND8 has a higher percentage of the longer poly(A) tails in the bloodstream form compared to the procyclic form [46]. COIII, CYb, C011, and COI on the other hand have a higher ratio of long poly(A) tails in the procyclic form [42, 46]. ND4 and MURFI have a similar size distribution between the two forms [46] (Table 2). Poly(A) tail addition in prokarya, such as Escherichia coll; acts as a targeting signal for mRNA decay. In the cytosol of eukarya, including humans and Saccharomyces cerevisiae, poly(A) tail addition protects them from mRNA degradation. In the mitochondria of S. cerevisiae, the poly(A) tail addition is 17 dispensable without effects on the mRNA metabolism. Plant mitochondrial and chloroplast mRNAs are also not constitutively polyadenylated, although there is evidence that a small proportion of each mRNA can be polyadenylated. Recent data have shown that the polyadenylation in plant mitochondria and chloroplasts can actually trigger degradation of the mRNA [54-57]. Human H-strand mitochondrial mRNAs are all polyadenylated with 50-60 residues, generating functional stop codons and conferring mt-mRNA stability [58-60]. It seems that 72 brucei mitochondria is another decay system regulated partially by polyadenylation, but unlike other systems, the edited state also seems to regulate the mRNA decay. Regulation Summary. 7? brucei mitochondria show numerous levels of post transcriptional regulation. Alhough very little is known about mRNA processing, it is clear that this processing is unique and important in the overall mitochondrial gene expression. Therefore, identifying and learning about the machinery that performs these tasks is important. We believe the PPR proteins will play a important role, as described below. 18 Regulation of 1: brucei Mitochondrial Expression Higher Steady . Fully Involved In mRNA State mRNA Higher 200'” nt Edited/Higher Mitochondrial Mitochondrial poly(A) tail ratio Complex level Levels Respiration ND1 Constitutive Never Edited Bloodstream l N03 Bloodstream Bloodstream Bloodstream Bloodstream l ND4 Bloodstream Similar Never Edited Bloodstream l N05 Bloodstream Single size Never Edited Bloodstream l ND7 Bloodstream Bloodstream Bloodstream Bloodstream l NDB Bloodstream Bloodstream Bloodstream Bloodstream l N09 Bloodstream Constitutive Bloodstream l CYb Procyclic Procyclic Procyclic Procyclic lll COI Procyclic Procyclic Never Edited Procyclic IV COll Procyclic Procyclic Procyclic Procyclic IV COlll Procyclic Procyclic lV A6 Constitutive Constitutive Both V MURFI Bloodstream Similar Never Edited MURFII Constitutive CR3 CR4 Bloodstream RSP12 Bloodstream Bloodstream Ribosomal Table 2. Regulation of 72 brucei mitochondrial gene expression. For each mitochondrial mRNA it is listed which lifecycle form has the higher steady state mRNA level, the higher 200nt:20nt poly(A) tail ratio, the higher level of fully edited mRNA, the lifecycle form in which it is thought to function, and in the mitochondrial complex in which it acts. TABLE 2 19 PPR Proteins Pentatricopeptide Repeat (PPR) proteins contain multiple, tandem PPR motifs that are degenerate 35-amino acid sequences that form two antiparallel a—helices with characteristic distributions of hydrophobic and hydrophilic amino acids. Multiple PPR domains are thought to form a superhelix with a central groove proposed to serve as the ligand-binding surface. The width of the groove is sufficient to hold a single strand of RNA. The sidechains lining the central groove are almost all hydrophobic, with positive residues at the bottom to bind the phosphate backbone. This suggests that the PPR motifs may be RNA-binding instead of protein-binding. The PPR proteins have been implicated in stability, translation, mRNA editing and mRNA processing of mitochondrial and chloroplast encoded messages. The largest groups of PPR proteins are found in the higher plants like Arabidopsis, rice, and maize probably due to the complex organelle gene expression in these systems. In all studies with PPR proteins thus far, mutants were not rescued by the other PPR proteins present in the cell, indicating that although there are many PPR proteins, they are not completely redundant. Pet309 and cya5 are PPR proteins implicated in stability and translation of mitochondrial cytochrome c oxidase subunit 1 (COXI) transcripts in yeast and neurospora respectively [61-64]. Maize CRP1 is required for translation of plastid petA and petD transcripts and in processing of petD from the polycistronic 20 precursor in the chloroplast [65, 66]. There are cytoplasmic male sterility (CMS) restorer genes that encode PPR proteins in petunia (Rf1)[67], radish (Rfk and Rfo)[68-70], as well as rice (Rf-1). Rf-1 is responsible for processing atp6 mRNA from polycistronic precursor [71, 72]. Most of the other PPR CMS restorers are involved through mRNA stability or processing as well [1]. Arabidopsis HCF152 null mutant shows impaired 5’-end processing and splicing of petB transcripts and the HCF152 protein has been found to bind the exon-intron junction of this RNA with high affinity [73-75]. In higher eukaryotes, Drosophila melanogaster BSF PPR protein binds a region of the bicoidmRNA 3’ untranslated region that supports normal mRNA maternal deposition and localization in the embryo during oogenesis through its role in mRNA stabilization. Drosophi/a BSF PPR protein is the only one studied to date that does not have an organellar localization [76]. A single missense mutation in the human LRP130/LRPPC gene, encoding the only human PPR protein studied, is the cause of the genetic LSFC disease, Leigh syndrome French Canadian, characterized by COX1 deficiency [77]. Levels of COXI and COXIII mRNAs were measurably reduced in LSFC patients. Translation of COXI is also reduced, indicating a role in translation or stability of mRNA for COXI [78]. The LRPPC protein has also been found to associate with mRNA/mRNP complexes [79-81] (Table 3). These studies show the involvement of PPR proteins in different aspects of organellar biogenesis. Several of the processes affected in plants are similar in 72 brucel; such as processing of polycistronic messages. 21 There have been a few PPR proteins studied that do not fit in to the umbrella description of being organellar targeted and involved in the processes of mitochondrial mRNA stability, processing and translation. There has only been one PPR protein studied to date that is involved in mRNA editing. Mutations in the ch4 gene of Arabidopsis thaliana have been found to be responsible for a decrease in the level of the NA(P)DH dehydrogenase (NDH) complex. There is no alteration in the size or level of the NDH transcripts, but by restriction analysis the mutants were found to be defective in one of their C to U editing sites in the initiation codon of the nth message [82]. The authors concluded that the PPR mutants are unable to edit this necessary NDH complex message [82] (Table 3). Researchers studying other PPR proteins in plants have looked for effects on RNA editing, but have not found any others as of yet. This may be due to editing mutations not presenting a obvious phenotype. This also may be due to the enormity of possible editing sites within the plant organellar genomes, making it hard to scan for these mutations. In the future, many other PPR proteins will probably be found to be involved in plant organellar RNA editing. 22 Eukaryotic PPR Proteins £r_g_anism PPR protein Gene Affected Affect Yeast Pet309 COXl Stability & Translation Neurospora Cyas COX! Stability 8 Translation Human LRP130 COXl Stability & Translation ? Drosophila BSF bicoid Stability Maize Crp1 petA & psaC 8 petD Translation petB/petD Processing Rice Rf-1 atp6 Processingfi Arabidopsis HCF 152 psbB-pst-pst-petB-petD Processing Wheat p63 COXII Transcription Initiation Arabidopsis CRR4 nth mRNA editing Table 3. Eukaryotic PPR proteins. The table lists representative PPR proteins that have been studied as well as their parent organism, what genes/mRNAs they affect and which mRNA process they affect. TABLE 3 23 There are many PPR proteins that may bind mRNA, but there is little direct evidence of this activity. Radish p67 and Drosophila BSF were both purified as sequence-specific RNA binding proteins in vitro [66, 76]. The C- terminal half of LRP130 has been shown to bind single-stranded polyadenylated RNA in vivo [83, 84]. Arabidopsis thaliana Hcf152 is a chloroplast protein that binds as a homodimer to the petB exon-intron junctions with high affinity [73- 75]. The strongest evidence comes from recent immunoprecipitations and microarray analysis showing that CRP1 binds specifically to petA and psaC mitochondrial mRNA messages in maize [85]. The same group that identified over 400 PPR proteins in plants have identified 18 PPR proteins in the 72 brucei database [86](T able 4). Next to plants, Trypanosomes have the most PPR proteins of all the organisms investigated. This may be relevant because during evolution Trypanosomes may have had a secondary loss of chloroplasts and they still contain some chloroplast- specific metabolic proteins today [87, 88]. They also both have complex organellar mRNA processing, including polycistronic precursor processing and mRNA editing. We have initiated a project to identify and characterize PPR proteins within I brace/2 We then proceeded to investigate the role of one of these proteins, TbPPR1, in 72 brucei mitochondrial biology. TbPPR1 was found to influence the growth rate of the 72 brucei cells and affect the expression of several mitochondrial messages. 24 PPR proteins are specific to Eukaryotes Organism Genes PPR Hits Homo sapiens 37,490 6 Drosophila melanogaster 17,087 2 Caenorhabditis elegans 20,673 2 Schizosacrhammyces pombe 5,010 6 Saccharomyces cerewlsrae 6,304 5 Trypanosoma brucei 16,757 19 CYan/dioschyzon merolae 4,772 10 Arabidopsis thaliana 28,581 470 Oryza sativa 74,385 655 Ralstonia solanacearum 5,118 1 Ric/rettsia pmwazekii 834 0 Synechocystls 5p. 3, 169 0 Table 4. PPR Proteins are Specific to Eukaryotes. The number of PPR genes (PPR Hits) found through bioinformatics in the eukaryotic organisms that have been sequenced is shown. Genes=number of identified genes for each organism. Modified from: Lurin, C., et al., Genome-wide analysis of Arabidopsis pentatricopeptide repeat proteins reveals their essential role in organelle biogenesis. Plant Cell, 2004. 16(8): p. 2089-103. TABLE 4 25 Overview A new family of over 400 proteins with PPR domains have been identified, but most of which are in plants. All of the investigated proteins in this family have been implicated in RNA metabolism in the mitochondria and chloroplasts, with the exception of Drosophila melanogaster BSF. The massive expansion of this protein family in the plant kingdom is most likely due to the high complexity of organellar expression in plants, including polycistronic processing and RNA editing. 72 bruceialso undergoes complex mitochondrial RNA processing to obtain a complete message. The mRNAs must be excised out from the polycistronic precursor, some are edited, and all of them must be polyadenylated before translation can occur. There is a great deal of information on 72 brucei mRNA editing and research on polyadenylation is progressing, but nothing is known about the processing of these messages from the polycistronic precursors. I have performed bioinformatics to characterize this family of proteins in 72 brucel; which, next to plants, has the highest number of PPR proteins of the sequenced eukaryotes. The majority of the 23 PPR proteins identified in 72 brucei have the traditional mitochondrial targeting signals. Knockdown of one of these proteins produces a slow growth phenotype and alterations in mitochondrial mRNA expression. Based on the current PPR literature and our own research, this family of proteins is highly likely to be involved in 72 hrucei mitochondrial mRNA expression. 26 My hypothesis is that PPR proteins in 72 brucei are involved in mitochondrial mRNA metabolism through processing of the polycistronic maxicircle transcript and/or mRNA stabilization of the processed products or mRNA translation. With this study, we have started to characterize the role these important proteins play in 72 hrucei mitochondrial biogenesis. 27 CHAPTER 2 BIOINFORMATICS 28 Introduction Trypanosoma brucei is a parasitic protozoan with a very complex lifecycle. Throughout the lifecycle the metabolism and the mitochondria of the protozoan go through many changes. The factors that control this complex regulation of metabolism and mitochondrial expression between the two lifecycle stages is unknown, but may be related to a family of Pentatricopeptide Repeat (PPR) proteins present in other organisms and 72 brucei itself. Through studies in plants[1, 2, 65-75, 82, 85, 86, 89-98], humans [76-81, 83, 84, 99], Drosophila [76], Chlamydomonas reinhardtii [100], yeast [61-63], and Neurosopora crassa [64, 101], the PPR proteins are emerging as important factors in organellar mRNA metabolism. - It was not until the Arabidopsis sequencing project, that the enormity of this family, over 400 putative PPR proteins, within Arabidopsis was realized. Almost all of the PPR proteins studied are implicated in organellar RNA metabolism. Along with all the PPR proteins identified in Arabidopsis, Lurin et. al. [86] identified 18 PPR proteins within the 72 brucei genome. This is the highest number of PPR proteins identified in non-plant eukaryotes. We began looking at the 18 putative 72 hrucei PPR proteins through bioinformatics. Of these 18 proteins, 13 have conventional mitochondrial targeting signals predicted by both Predotar (http://genoplante—info.infobiogen.fr/predotar/) and MitoProt [102], suggesting that this class of proteins plays an important role in Trypanosome mitochondrial biology. We identified the PPR domains in the 18 29 putative proteins through Pfam and created a probability matrix of the motif in HMMER and MEME. Using this T. bruceispeciflc probability matrix, we identified 6 other putative PPR proteins within 72 brace/2 and additional PPR domains were found in almost all of the PPR proteins previously identified in Lurin et al. [86]. Materials and Methods Protein and Motif Identification. The PPR domains from the 18 identified proteins [86] were found with Pfam [103] and aligned in ClustalX version 1.81 [104]. The ClustalX alignment was imported into both the HMMER 2.2 package (http://hmmer.wustl.edu/) [105, 106] and MEME [107]. Using hmmbuild and hmmcalibrate, HMMER constructed a probability matrix or position-specific scoring system of the 72 brucei PPR motif using profile hidden Markov models which can put in additions and deletions anywhere within the alignment. This was then used in hmmsearch for a more extensive search of the 72 brace/2 Leishmania major, and Trypanosoma cruzi databases for other PPR proteins. MEME [107] used a slightly different algorithm, two-component finite mixture model, which uses a “motif” rather then “profile" method to produce alignments of ungapped blocks. MEME also gave us a slightly more 72 brucei specific motif to search the databases with (Fig. 5). All amino acids in the MEME have a probability of 0.2 or higher, with the highest probability amino acids listed on the first line. For the consensus sequence from the HMMER the capital letters have probability >05 of occurrence. 30 Mitmhonorial Targeting. Mitochondrial targeting was determined through TargetP [108], Predotar (http://genoplante-info.infobiogen.fr/predotar/), and MitoProt [102]. Dogma; We also used the following databases in our search for kinetoplast and other eukaryotic Orthologs: Sanger www.sanger.ac.uk 1" a.“ TIGR www.tigr.org Trypanosoma cruzi Database tcruzidb.org Results PPR nen i i nifl w'hHMMER n ME We began by formulating a 72 brucei specific consensus PPR motif and comparing it to the Arabidopsis to see if we could find species specific differences. A Clustal X alignment of all the Pfam identified PPR domains within the 18 PPR proteins identified by Lurin et al. [86] was imported into both HMMER and MEME. HMMER then calculated a probability matrix based on profile hidden Markov models for the consensus sequence. Capital letters within the consensus indicate a probability of occurrence >0.5. For 16 out of the 35 positions, the 72 brucei PPR HMMER consensus sequence was identical to the Arabidopsis PPR consensus sequence (Fig. 5). The only differences between the two sequences were conservative replacements such as leucine in position 6 in place of 31 isoleucine found in Arabidopsis, arginine at position 8 in place of asparagine, cysteine at position 10 in place of tyrosine, and arginine at position 12 in place of lysine. The 72 brucei HMMER consensus included additional amino acids where the Arabidopsis consensus did not list any, including an alanine at position 13, an aspartic acid at position 15, tryptophan at position 16, glutamic acids at 20, 21, 24, 25, 28 and 36, arginines at positions 27 and 29, a valine at position 31, and proline at 32 and 35. These additions were probably based on the smaller .1; 1. mm 1.3-- .“ number of PPR domains aligned in 72 brucei versus the enormous number of Arabidopsis PPR domains that were used for the same type of consensus sequence. All of these additions do provide a consensus sequence that is more specific for 72 brucel; but with only 5 differences between the two sequences being conservative replacements and 16 out of the 35 amino acids being identical, this gives us a highly conserved PPR motif (Fig. 5). The Clustal X alignment of PPR domains was also imported into MEME [107] for a consensus sequence based on the two-component finite mixture model. The T. brucei MEME consensus included amino acids above a probability of occurrence at 0.2. The only differences the MEME and HMMER programs found between the T. brucei consensus sequences were at positions 21 and 23, with the tyrosine to leucine change at position 23 being a conservative replacement (Fig. 5). 32 7'. brucei Specific PPR Motif PPR COIISCTISUS . 123456789flflQflfifififlflflmz28253288$32$$$$ T-bmce‘ .mxlt xlflcfll‘k A131) ..... AIYEEIERERIEVxL—gfi]. MEME T I 1. v L Q K D Y F T. brucei human 1 rgcgt aQ'd ew- eml )e eEr e (EB-v pmp e HMMER T HelixA HelixB Arabidopsis .nm1~myx._ .-. .ly. m. . .I- .nl. thalianaa [:Iomsewedbetmenbothorgarisrrs Trypanosomabmoeispeoific Figure 5. 72 brucei specific PPR domains. Helix A and Helix B denote the 1.! . ‘4. _‘ area where the two a—helices of the PPR motif are predicted to be. The MEME consensus sequence is listed first with alternative amino acids on the second and third line. All amino acids in the MEME have a probability of 0.2 or higher, with the highest probability amino acids listed on the first line. The consensus sequence generated from the HMMER program is aligned underneath the MEME with the capital letters having a probability of >0.5. The Arabidopsis thaliana PPR consensus sequence published by Small et al. [1] is shown last. Amino acids conserved between both organisms are boxed and the T. brucei specific amino acids found are FIGURE 5 33 Protein Characterization With the HMMER program [105, 106] we identified 14 additional PPR motifs on most (10 of 17) of the 72 brucei PPR proteins already identified by Lurin et al. [86]. (One of the PPR proteins found by Lurin et al. (Tb11.01.5980) was also dropped because the 2 PPR domains identified were not in tandem.) A representation of the conservation in the 72 brucei domains aligned with the HMMER 72 brucei and the Pfam PPR consensus sequences is shown in Figure 6, E .4 I". vases—5.51:1 using the Tb927.2.3180 PPR domains as representative domains. 34 Figure 6. TbPPR1 PPR motif sequences. Depleted here is an alignment of all the PPR motifs found in TbPPR1 as well as the Pfam consensus motif and the HMMER consensus motif. The light gray amino acids are the amino acids in the TbPPR1 motifs that match the HMMER consensus motif. The medium gray amino acids in the TbPPR1 motifs are what HMMER considers a conservative replacement to the HMMER consensus motif. The outlines boxes of the TbPPR1 motif sequences are amino acids that match the Pfam consensus motif sequence. The portions of the sequence corresponding to helix A or B are above the labeled cylinder for that helix. 3S TbPPR1 PPR motif sequences Position T.brucei HMMER Pfam PPR TbPPR1-1 TbPPR1-2 TbPPR1-3 TbPPR1-4 TbPPR1-5 TbPPR1-6 TbPPR1-7 TbPPR1-8 TbPPR1-9 TbPPR1-10 TbPPR1-11 TbPPR1-12 TbPPR1-13 TbPPR1-14 .5 ‘1 .s (D N 0 Position T.brucei HMMER 21 f .K-‘svasrly-cr-<-< 22 23 24 25 26 27 28 l 29 30 FT— V 31 32 33 w “on Pfam PPR TbPPR1-1 TbPPR1-2 TbPPR1-3 TbPPR1-4 TbPPR1-5 TbPPR1-6 TbPPR1-7 TbPPR1-8 TbPPR1-9 TbPPR1-10 TbPPR1-11 TbPPR1-12 TbPPR1-13 TbPPR1-14 {—l— ,-_--—L \uulmwm ——>u mm<>ommm>>m G) v -mmmm. l‘l'l 4mmm0m HMMER consensus HMMER conserved replacement "I Pfam consensus e e —zm_m mm . > .- VI" f E K _' Q 36 e E r—‘Dm'z‘r'omlm mg. 33-<‘:; 13 FIGURE 6 .HUifr k - --l>€‘!""‘ F (DI 4:013» 1:; '1 D X 93> m my. all!" w 74nt Unedited CYb 45nt Primer """ Ratio 0.09 0.08 0.01 0.05 1.14 0.76 0.01 0.03 0.64 0.47 Edited/Unedited FIGURE 17 75 Discussion and Conclusion The family of PPR proteins is an interesting and greatly expanding family in plants. Therefore their functions in 72 brucei should also prove to be of great interest. Based on analysis of the other RNA binding proteins, PPR proteins show a high probability of being RNA binding proteins. Studies of plant PPR proteins and those in other eukaryotes, show that they are involved in different processes of organellar mRNA expression. Our bioinformatics studies built on the 18 72 brucei PPR proteins found by Lurin et al. [86], expanding the family to 23 proteins in 72 brucei and finding more PPR domains in the proteins already discovered. It also showed that this is a family of proteins that is well conserved throughout the kinetoplastid family. Knocking down TbPPR1 (Tb927.2.3180) through RNAi has a clear effect on the growth of 72 brucei. This same growth defect was seen using this plasmid for RNAi in other studies [117-121]. The slow growth phenotype that accompanied the knockdown of TbPPR1 along with the importance of PPR proteins in other organisms led us to further investigate the function of TbPPR1 in the mitochondrial RNA metabolism in 72 brucei. The ribosomal RNAs decrease in expression and ND1, ND7, ND8 and ND9 mRNAs all remain relatively unchanged during TbPPR1 RNAi induction. The ribosomal decrease may be a direct or indirect effect from the loss of TbPPR1, but not enough is understood about 72 brucei mitochondrial biogenesis to know for sure. A similar ribosomal 76 RNA decrease also occurs in the cytoplasm of bacteria when undergoing stress induced by pH changes [122]. For CYb mRNA, the amount of long poly(A) tail size class present during TbPPR1 RNAi induction, seems to decrease whereas the short poly(A) tail size class is stabilized. ND4 also appeared to lose the long poly(A) tail size class, but without any stabilization of the short poly(A) tail size class; however quantification showed that there is a decrease in both poly(A) tail size classes. There are several possibilities to explain why the long poly(A) tail size class of CYb decreased. One theory is that there may be a defect in polyadenylation. Loss of TbPPR1 could be affecting the poly(A) machinery itself or a simple loss of ATP to incorporate into the poly(A) tails due to cellular stress as the TbPPR1 RNAi progresses. It could also be due to loss of CYb editing as shown in the poison primer extension analysis. It has been shown that the majority of the long poly(A) tail size class is composed of edited CYb messages, where the unedited messages are in the short poly(A) tail size class [43]. We do know that for the RP512 mRNA, the edited form is targeted for destruction when it does not posses a poly(A) tail. In contrast the unedited form is targeted for destruction when it does contain a poly(A) tail [50, 123]. So for this mRNA, there is a connection between the editing state, poly(A) tail addition and mRNA degradation which may also be occurring for CYb. Both mRNA editing and processing/polyadenylation have been shown to be two independent events [31]. Based on this information, the composition of the CYb poly(A) tails in the TbPPR1 77 knockdowns should be studied next. The loss of ND4 would not fit this hypothesis since ND4 is never edited. Unfortunately, there is not yet enough information available on T. brucei mitochondria RNA metabolism to speculate any further. Further studies should then tell us if transcripts within the bloodstream form are affected by TbPPR1 RNAi and whether the protein expression is affected confirming a role in mitochondrial mRNA biogenesis. Further structural studies should also confirm that these PPR proteins are in fact RNA binding proteins. Future Work We have been attempting to confirm the Northern results using Real Time PCR, but we are still optimizing this procedure. We have developed good primer pairs that produce high quality melt curves for almost every mitochondrial transcript. Primer concentrations and cDNA concentrations are still being optimized for each primer set. We have also begun to characterize the poly(A) tails in the TbPPR1 RNAi cells through 5’ and 3’ RACE, but more primer pair optimization must also be worked out here. Regulation of both the TbPPR1 protein and RNA from both bloodstream and procyclic Trypanosomes can be further studied with Westerns and Northerns respectively. Many characterization studies can also be done when we in vivo epitope label the PPR1 protein or obtain antibodies against the TbPPR1 protein. Western blot protein characterization, immunolocalization, subcellular 78 fractionation, and protein complex studies can all be done with these epitope tags and/or antibodies. We are also working on a collaboration to express the TbPPR proteins and perform crystallography and NMR studies with the proteins and their RNA substrates (determined through mRNA expression alterations via Northern blots and ribonuclease protection assays) to look at the binding surface of the PPR motif. These proteins and RNA substrates will also be used in gel mobility shift assays to determine the kinetics of binding. Eventually we hope to use several different approaches to find some drugs that target these PPR proteins. 79 BIBLIOGRAPHY 8O 10. ll. 12. 13. 14. Small, ID. and N. Peeters, The PPR motif - a TPR-related motif prevalent in plant organellar proteins. Trends Biochem Sci, 2000. 25(2): p. 46-7. Williams, RM. and A. Barkan, A chloroplast-localized PPR protein required for plastid ribosome accumulation. Plant J, 2003. 36(5): p. 675-86. Pepin, J., et al., Trial of prednisolone for prevention of melarsoprol-induced encephalopathy in gambiense sleeping sickness. Lancet, 1989. 1(8649): p. 1246- 50. Vickerman, K. and AG. Luckins, Localization of variable antigens in the surface coat of Trypanosoma brucei using ferritin conjugated antibody. Nature, 1969. 224(224): p. 1125—6. Le Ray, D., J .D. Barry, and K. Vickerrnan, Antigenic heterogeneity of metacyclic forms of Trypanosoma brucei. Nature, 1978. 273(5660): p. 300-2. Jenni, L., Comparisons of antigenic types of Trypanosoma (T)brucei strains transmitted by Glossina m. morsitans. Acta Trop, 1977. 34(1): p. 35-41. Ryley, J .F., Studies on the metabolism of the protozoa. 9. Comparative metabolism of blood-stream and culture forms of Trypanosoma rhodesiense. Biochem J, 1962. 85: p. 211-23. Flynn, I.W. and LB. Bowman, The metabolism of carbohydrate by pleomorphic African trypanosomes. Comp Biochem Physiol B, 1973. 45(1): p. 25-42. Cazzulo, J .J ., Aerobic fermentation of glucose by trypanosomatids. Faseb J, 1992. 6(13): p. 3153-61. Lamour, N., et al., Praline metabolism in procyclic Trypanosoma brucei is down- regulated in the presence of glucose. J Biol Chem, 2005. 280(12): p. 11902-10. Grant, RT. and J .R. Sargent, L-alpha-Glycerophosphate dehydrogenase, a component of an oxidase system in Trypanosoma rhodesiense. Biochem J, 1961. 81: p. 206-14. Bienen, E.J., E. Hammadi, and CC. Hill, Trypanosoma brucei: biochemical and morphological changes during in vitro transformation of bloodstream- to procyclic-trypomastigotes. Exp Parasitol, 1981. 51(3): p. 408-17. Beattie, D.S. and M.M. Howton, The presence of rotenone-sensitive NADH dehydrogenase in the long slender bloodstream and the procyclicforms of Trypanosoma brucei brucei. Eur J Biochem, 1996. 241(3): p. 888-94. Nolan, DP. and HF. Voorheis, The mitochondrion in bloodstream forms of Trypanosoma brucei is energized by the electrogenic pumping of protons catalysed by the F I F O-ATPase. Eur J Biochem, 1992. 209(1): p. 207-16. 81 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. Bochud-Allemann, N. and A. Schneider, Mitochondrial substrate level phosphorylation is essential for growth of procyclic Trypanosoma brucei. J Biol Chem, 2002. 277(36): p. 32849-54. Coustou, V., et al., ATP generation in the Trypanosoma brucei procyclic form: cytosolic substrate level is essential, but not oxidative phosphorylation. J Biol Chem, 2003. 278(49): p. 49625-35. van Weelden, S.W., et al., Procyclic Trypanosoma brucei do not use Krebs cycle activity for energy generation. J Biol Chem, 2003. 278(15): p. 12854-63. Schultz, BE. and SI. Chan, Structures and proton-pumping strategies of mitochondrial respiratory enzymes. Annu Rev Biophys Biomol Struct, 2001. 30: p. 23-65. Walker, R., Jr., et al., The effect of over-expression of the alternative oxidase in the procyclic forms of Trypanosoma brucei. Mol Biochem Parasitol, 2005. 139(2): p. 153-62. Fang, J. and D.S. Beattie, Alternative oxidase present in procyclic Trypanosoma brucei may act to lower the mitochondrial production of superoxide. Arch Biochem Biophys, 2003. 414(2): p. 294-302. Michelotti, BF. and S.L. Hajduk, Developmental regulation of trypanosome mitochondrial gene expression. J Biol Chem, 1987. 262(2): p. 927-32. Michelotti, E.F., et al., Trypanosoma brucei mitochondrial ribosomal RNA synthesis, processing and developmentally regulated expression. Mol Biochem Parasitol, 1992. 54(1): p. 31-41. Myler, P.J., et al., Structural organization of the maxicircle variable region of Trypanosoma brucei: identification of potential replication origins and topoisomerase [1 binding sites. Nucleic Acids Res, 1993. 21(3): p. 687-94. Clement, S.L., M.K. Mingler, and DJ. Koslowsky, An Intragenic Guide RNA Location Suggests a Complex Mechanism for Mitochondrial Gene Expression in Trypanosoma brucei. Eukaryot Cell, 2004. 3(4): p. 862-9. J asmer, DP. and K. Stuart, Conservation of kinetoplastid minicircle characteristics without nucleotide sequence conservation. Mol Biochem Parasitol, 1986. 18(3): p. 257-69. Pollard, V.W. and S.L. Hajduk, Trypanosoma equiperdum Minicircles Encode Three Distinct Primary Transcripts Which Exhibit Guide RNA Characteristics. molecular and Cellular Biology, 1991: p. 1668-1675. 82 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. Koslowsky, D.J., et al., Guide RNAs for transcripts with developmentally regulated RNA editing are present in both life cycle stages of Trypanosoma brucei. Mol Cell Biol, 1992. 12(5): p. 2043-9. Hong, M. and L. Simpson, Genomic organization of Trypanosoma brucei kinetoplast DNA minicircles. Protist, 2003. 154(2): p. 265-79. Riley, G.R., R.A. Corell, and K. Stuart, Multiple guide RNAs for identical editing of Trypanosoma brucei apocytochrome b mRNA have an unusual minicircle location and are developmentally regulated. J Biol Chem, 1994. 269(8): p. 6101- 8. Grams, J ., M.T. McManus, and S.L. Hajduk, Processing of polycistronic guide RNAs is associated with RNA editing complexes in Trypanosoma brucei. The EMBO Journal, 2000. 19(20): p. 5525-5532. Koslowsky, DJ. and G. Yahampath, Mitochondrial mRNA 3' cleavage/polyadenylation and RNA editing in Trypanosoma brucei are independent events. Mol Biochem Parasitol, 1997. 90(1): p. 81-94. Read, L.K., P.J. Myler, and K. Stuart, Extensive editing of both processed and preprocessed maxicircle CR6 transcripts in Trypanosoma brucei. J Biol Chem, 1992. 267(2): p. 1123-8. Feagin, J .E., D.P. J asmer, and K. Stuart, Apocytochrome b and other mitochondrial DNA sequences are differentially expressed during the life cycle of Trypanosoma brucei. Nucleic Acids Res, 1985. 13(12): p. 4577-96. J asmer, D.P., J .E. Feagin, and K. Stuart, Diverse patterns of expression of the cytochrome c oxidase subunit I gene and unassigned reading frames 4 and 5 during the life cycle of Trypanosoma brucei. Mol Cell Biol, 1985. 5(11): p. 3041- 7. Feagin, J .E. and K. Stuart, Diflerential expression of mitochondrial genes between life cycle stages of Trypanosoma brucei. Proc Natl Acad Sci U S A, 1985. 82(10): p. 3380-4. Feagin, J .E., D.P. J asmer, and K. Stuart, Developmentally regulated addition of nucleotides within apocytochrome b transcripts in Trypanosoma brucei. Cell, 1987. 49(3): p. 337-45. Souza, A.E., P.J. Myler, and K. Stuart, Maxicircle CR1 transcripts of Trypanosoma brucei are edited and developmentally regulated and encode a putative iron-sulfur protein homologous to an NADH dehydrogenase subunit. Mol Cell Biol, 1992. 12(5): p. 2100-7. 83 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. Souza, A.E., et al., Extensive editing of CR2 maxicircle transcripts of Trypanosoma brucei predicts a protein with homology to a subunit of NADH dehydrogenase. Mol Cell Biol, 1993. 13(11): p. 6832-40. Koslowsky, D.J., et al., The M URF 3 gene of T. brucei contains multiple domains of extensive editing and is homologous to a subunit of NADH dehydrogenase. Cell, 1990. 62(5): p. 901-11. Corell, R.A., P. Myler, and K. Stuart, Trypanosoma brucei mitochondrial CR4 gene encodes an extensively edited mRNA with completely edited sequence only in bloodstream forms. Mol Biochem Parasitol, 1994. 64(1): p. 65-74. Priest, J .W. and S.L. Hajduk, Developmental regulation of Mitochondrial Biogenesis in Trypanosoma brucei. Journal of Bioenergetics and Biomemebranes, 1994. 26(2): p. 179-191. Feagin, J .E., J .M. Abraham, and K. Stuart, Extensive editing of the cytochrome c oxidase III transcript in Trypanosoma brucei. Cell, 1988. 53(3): p. 413—22. Feagin, J .E. and K. Stuart, Developmental aspects of uridine addition within mitochondrial transcripts of Trypanosoma brucei. Mol Cell Biol, 1988. 8(3): p. 1259-65. - Blum, B., N. Bakalara, and L. Simpson, A model for RNA editing in kinetoplastid mitochondria: "guide" RNA molecules transcribed from maxicircle DNA provide the edited information. Cell, 1990. 60(2): p. 189-98. Bhat, G.J., et al., An extensively edited mitochondrial transcript in kinetoplastids encodes a protein homologous to ATPase subunit 6. Cell, 1990. 61(5): p. 885-94. Bhat, G.J., et al., Transcript-specific developmental regulation of polyadenylation in Trypanosoma brucei mitochondria. Mol Biochem Parasitol, 1992. 52(2): p. 231-40. Militello, K.T. and L.K. Read, Coordination of kRNA editing and polyadenylation in Trypanosoma brucei mitochondria: complete editing is not required for long poly(A) tract addition. Nucleic Acids Res, 1999. 27(5): p. 1377-85. Read, L.K., et al., Editing of Trypanosoma brucei maxicircle CR5 mRNA generates variable carboxy terminal predicted protein sequences. Nucleic Acids Res, 1994. 22(8): p. 1489-95. Feagin, J .E., D.P. J asmer, and K. Stuart, Differential mitochondrial gene expression between slender and stumpy bloodforms of Trypanosoma brucei. Mol Biochem Parasitol, 1986. 20(3): p. 207-14. 84 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. Kao, CY. and L.K. Read, Opposing efi‘ects of polyadenylation on the stability of edited and unedited mitochondrial RNAs in Trypanosoma brucei. Mol Cell Biol, 2005. 25(5): p. 1634-44. Militello, K.T. and L.K. Read, UTP-dependent and -independent pathways of mRNA turnover in Trypanosoma brucei mitochondria. Mol Cell Biol, 2000. 20(7): p. 2308-16. Decker, C]. and B. Sollner—Webb, RNA editing involves indiscriminate U changes throughout precisely defined editing domains. Cell, 1990. 61(6): p. 1001- 11. Read, L.K., et al., Developmental regulation of RNA editing and polyadenylation in four life cycle stages of Trypanosoma congolense. Mol Biochem Parasitol, 1994. 68(2): p. 297-306. Gagliardi, D. and C]. Leaver, Polyadenylation accelerates the degradation of the mitochondrial mRNA associated with cytoplasmic male sterility in sunflower. Embo J, 1999. 18(13): p. 3757-66. Gagliardi, D., et al., Plant mitochondrial polyadenylated mRNAs are degraded by a 3 '- to 5 '-exoribonuclease activity, which proceeds unimpeded by stable secondary structures. J Biol Chem, 2001. 276(47): p. 43541-7. Lupold, D.S., A.G. Caoile, and DB. Stern, Polyadenylation occurs at multiple sites in maize mitochondrial cox2 mRNA and is independent of editing status. Plant Cell, 1999. 11(8): p. 1565-78. Kuhn, J ., U. Tengler, and S. Binder, Transcript lifetime is balanced between stabilizing stem-loop structures and degradation-promoting polyadenylation in plant mitochondria. Mol Cell Biol, 2001. 21(3): p. 731-42. Ojala, D., J. Montoya, and G. Attardi, tRNA punctuation model of RNA processing in human mitochondria. Nature, 1981. 290(5806): p. 470-4. Temperley, R.J., et al., Investigation of a pathogenic mtDNA microdeletion reveals a translation-dependent deadenylation decay pathway in human mitochondria. Hum Mol Genet, 2003. 12(18): p. 2341-8. Gagliardi, D., et al., Messenger RNA stability in mitochondria: different means to an end. Trends Genet, 2004. 20(6): p. 260-7. Manthey, G.M., B.D. Przybyla-Zawislak, and J .E. McEwen, The Saccharomyces cerevisiae Pet309 protein is embedded in the mitochondrial inner membrane. Eur J Biochem, 1998. 255(1): p. 156-61. 85 62. 63. 65. 66. 67. 68. 69. 70. 71. 72. 73. Krause, K., et al., The mitochondrial message-specific mRNA protectors Cpr and Pet309 are associated in a high-molecular weight complex. Mol Biol Cell, 2004. 15(6): p. 2674-83. N aithani, S., et al., Interactions among COX1, C 0X2, and C 0X3 mRNA-specific translational activator proteins on the inner surface of the mitochondrial inner membrane of Saccharomyces cerevisiae. Mol Biol Cell, 2003. 14(1): p. 324-33. Coffin, J .W., et al., The Neurospora crassa cya-5 nuclear gene encodes a protein with a region of homology to the Saccharomyces cerevisiae PET309 protein and is required in a post-transcriptional step for the expression of the mitochondrially encoded COX1 protein. Curr Genet, 1997. 32(4): p. 273-80. Fisk, D.G., M.B. Walker, and A. Barkan, Molecular cloning of the maize gene crpI reveals similarity between regulators of mitochondrial and chloroplast gene expression. Embo J, 1999. 18(9): p. 2621-30. Lahmy, S., et al., A chloroplastic RNA-binding protein is a new member of the PPR family. FEBS Lett, 2000. 480(2-3): p. 255-60. Bentolila, S., A.A. Alfonso, and MR. Hanson, A pentatricopeptide repeat- containing gene restores fertility to cytoplasmic male-sterile plants. Proc N at] Acad Sci U S A, 2002. 99(16): p. 10887-92. Brown, G.G., et al., The radish Rfo restorer gene of Ogura cytoplasmic male sterility encodes a protein with multiple pentatricopeptide repeats. Plant J, 2003. 35(2): p. 262-72. Desloire, S., et al., Identification of the fertility restoration locus, Rfo, in radish, as a member of the pentatricopeptide-repeat protein family. EMBO Rep, 2003. 4(6): p. 588-94. Koizuka, N., et al., Genetic characterization of a pentatricopeptide repeat protein gene, orf68 7, that restores fertility in the cytoplasmic male-sterile Kosena radish. Plant J, 2003. 34(4): p. 407-15. Kazama, T. and K. Toriyama, A pentatricopeptide repeat-containing gene that promotes the processing of aberrant atp6 RNA of cytoplasmic male-sterile rice. FEBS Lett, 2003. 544(1-3): p. 99-102. Komori, T., et al., Map-based cloning of a fertility restorer gene, Rf-l, in rice (Oryza sativa L. ). Plant J, 2004. 37(3): p. 315-25. N akamura, T., et al., Chloroplast RNA-binding and pentatricopeptide repeat proteins. Biochem Soc Trans, 2004. 32(Pt 4): p. 571-4. 86 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. 84. 85. Nakamura, T., et al., RNA-binding properties of H CF 1 52, an Arabidopsis PPR protein involved in the processing of chloroplast RNA. Eur J Biochem, 2003. 270(20): p. 4070-81. Meierhoff, K., et al., HCF152, an Arabidopsis RNA binding pentatricopeptide repeat protein involved in the processing of chloroplast psbB-pst-pst—petB- petD RNAs. Plant Cell, 2003. 15(6): p. 1480-95. Mancebo, R., et al., BSF binds specifically to the bicoid mRNA 3 ' untranslated region and contributes to stabilization of bicoid mRNA. Mol Cell Biol, 2001. 21(10): p. 3462-71. Mootha, V.K., et al., Identification of a gene causing human cytochrome c oxidase deficiency by integrative genomics. Proc Natl Acad Sci U S A, 2003. 100(2): p. 605-10. Xu, F ., et al., The role of the LRPPRC (leucine-rich pentatricopeptide repeat cassette) gene in cytochrome oxidase assembly: mutation causes lowered levels of COX (cytochrome c oxidase) I and COX III mRNA. Biochem J, 2004. 382(Pt 1): p. 33 1-6. Ostrowski, J ., et al., Heterogeneous nuclear ribonucleoprotein K protein associates with multiple mitochondrial transcripts within the organelle. J Biol Chem, 2002. 277(8): p. 6303-10. Liu, L. and W.L. McKeehan, Sequence analysis of LRPPRC and its SEC 1 domain interaction partners suggests roles in cytoskeletal organization, vesicular trafficking, nucleocytosolic shuttling, and chromosome activity. Genomics, 2002. 79(1): p. 124-36. Mili, S., et al., Distinct RNP complexes of shuttling hnRNP proteins with pre- mRNA and mRNA: candidate intermediates in formation and export of mRNA. Mol Cell Biol, 2001. 21(21): p. 7307-19. Kotera, E., M. Tasaka, and T. Shikanai, A pentatricopeptide repeat protein is essential for RNA editing in chloroplasts. Nature, 2005. 433(7023): p. 326-30. Mili, S. and S. Pinol-Roma, LRP130, a pentatricopeptide motif protein with a noncanonical RNA-binding domain, is bound in vivo to mitochondrial and nuclear RNAs. Mol Cell Biol, 2003. 23(14): p. 4972-82. Tsuchiya, N., et al., LRP130, a single-stranded DNA/RNA-binding protein, localizes at the outer nuclear and endoplasmic reticulum membrane, and interacts with mRNA in vivo. Biochem Biophys Res Commun, 2004. 317(3): p. 736-43. Schmitz-Linneweber, C., R. Williams-Carrier, and A. Barkan, RNA Immunoprecipitation and Microarray Analysis Show a Chloroplast 87 86. 87. 88. 89. 90. 91. 92. 93. 94. 95. 96. 97. Pentatricopeptide Repeat Protein to Be Associated with the 5 ' Region of mRNAs Whose Translation It Activates. Plant Cell, 2005. Lurin, C., et al., Genome-wide analysis of Arabidopsis pentatricopeptide repeat proteins reveals their essential role in organelle biogenesis. Plant Cell, 2004. 16(8): p. 2089-103. Martin, W. and P. Borst, Secondary loss of chloroplasts in trypanosomes. Proc Natl Acad Sci U S A, 2003. 100(3): p. 765-7. Hannaert, V., et al., Plant-like traits associated with metabolism of Trypanosoma parasites. Proc Natl Acad Sci U S A, 2003. 100(3): p. 1067-71. Ikeda, T.M. and M.W. Gray, Characterization of a DNA-binding protein implicated in transcription in wheat mitochondria. M01 Cell Biol, 1999. 19(12): p. 8113-22. Oguchi, T., et al., Genomic structure of a novel Arabidopsis clock-controlled gene, AtC401, which encodes a pentatricopeptide repeat protein. Gene, 2004. 330: p. 29-37. Hashimoto, M., et al., A nucleus-encoded factor, CRR2, is essential for the expression of chloroplast nth in Arabidopsis. Plant J, 2003. 36(4): p. 541-9. Yamazaki, H., M. Tasaka, and T. Shikanai, PPR motifs of the nucleus-encoded factor, PGR3, function in the selective and distinct steps of chloroplast gene expression in Arabidopsis. Plant J, 2004. 38(1): p. 152-63. Cushing, D.A., et al., Arabidopsis embl 75 and other ppr knockout mutants reveal essential roles for pentatricopeptide repeat (PPR) proteins in plant embryogenesis. Planta, 2005. Barkan, A., et al., A nuclear mutation in maize blocks the processing and translation of several chloroplast mRNAs and provides evidence for the difierential translation of alternative mRNA forms. Embo J, 1994. 13(13): p. 3170-81. Gothandam, K.M., et al., OsPPRI, a pentatricopeptide repeat protein of rice is essential for the chloroplast biogenesis. Plant M01 Biol, 2005. 58(3): p. 421-33. Akagi, H., et al., Positional cloning of the rice Rf-I gene, a restorer of BT-type cytoplasmic male sterility that encodes a mitochondria-targeting PPR protein. Theor Appl Genet, 2004. 108(8): p. 1449-57. Klein, R.R., et al., Fertility restorer locus Rf] of sorghum (Sorghum bicolor L.) encodes a pentatricopeptide repeat protein not present in the colinear region of rice chromosome 12. Theor Appl Genet, 2005. 88 98. 99. 100. 101. 102. 103. 104. 105. 106. 107. 108. 109. 110. 111. Hattori, M., M. Hasebe, and M. Sugita, Identification and characterization of cDNAs encoding pentatricopeptide repeat proteins in the basal land plant, the moss Physcomitrella patens. Gene, 2004. 343(2): p. 305-11. Tsuchiya, N., et al., LRP130, a protein containing nine pentatricopeptide repeat motifs, interacts with a single-stranded cytosine-rich sequence of mouse hypervariable minisatellite Pc-I. Eur J Biochem, 2002. 269(12): p. 2927-33. Auchincloss, A.H., et al., Characterization of Tbc2, a nucleus-encoded factor specifically required for translation of the chloroplast pst mRNA in Chlamydomonas reinhardtii. J Cell Biol, 2002. 157(6): p. 953-62. Lown, F.J., A.T. Watson, and S. Purton, Chlamydomonas nuclear mutants that fail to assemble respiratory or photosynthetic electron transfer complexes. Biochem Soc Trans, 2001. 29(Pt 4): p. 452-5. Claros, MG. and P. Vincens, Computational method to predict mitochondrially imported proteins and their targeting sequences. Eur J Biochem, 1996. 241(3): p. 779-86. Bateman, A., et al., The Pfam protein families database. Nucleic Acids Res, 2004. 32(Database issue): p. D138-41. Thompson, J .D., et al., The CLUSTAL_X windows interface: flexible strategies for multiple sequence alignment aided by quality analysis tools. Nucleic Acids Res, 1997. 25(24): p. 4876-82. Eddy, S.R., Profile hidden Markov models. Bioinformatics, 1998. 14(9): p. 755- 63. Eddy, S.R., Hidden Markov models. Curr Opin Struct Biol, 1996. 6(3): p. 361-5. Bailey, TL. and C. Elkan, Fitting a mixture model by expectation maximization to discover motifs in biopolymers. Proc Int Conf Intell Syst M01 Biol, 1994. 2: p. 28- 36. Emanuelsson, 0., et al., Predicting subcellular localization of proteins based on their N-terminal amino acid sequence. J Mol Biol, 2000. 300(4): p. 1005-16. Treger, M. and E. Westhof, Statistical analysis of atomic contacts at RNA-protein interfaces. J M01 Recognit, 2001. 14(4): p. 199-214. Edwards, T.A., et al., Structure of Pumilio reveals similarity between RNA and peptide binding motifs. Cell, 2001. 105(2): p. 281-9. Rozen, S. and H. Skaletsky, Primer3 on the W for general users and for biologist programmers. Methods M01 Biol, 2000. 132: p. 365-86. 89 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. Altschul, S.F., et al., Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res, 1997. 25(17): p. 3389-402. Chomczynski, P. and N. Sacchi, Single—step method of RNA isolation by acid guanidinium thiocyanate-phenol—chloroform extraction. Anal Biochem, 1987. 162(1): p. 156-9. Motyka, SA. and RT. Englund, RNA interference for analysis of gene function in trypanosomatids. Curr Opin Microbiol, 2004. 7(4): p. 362-8. Shi, H., et al., Genetic interference in Trypanosoma brucei by heritable and inducible double-stranded RNA. Rna, 2000. 6(7): p. 1069-76. Djikeng, A., et al., RNA interference in Trypanosoma brucei: cloning of small interfering RNAs provides evidence for retroposon-derived 24-26-nucleotide RNAs. Rna, 2001. 7(11): p. 1522—30. Wang, Z., et al., Inhibition of Trypanosoma brucei gene expression by RNA interference using an integratable vector with opposing T7 promoters. J Biol Chem, 2000. 275(51): p. 40174-9. Liu, Q., et al., Identification and functional characterization of lsm proteins in Trypanosoma brucei. J Biol Chem, 2004. 279(18): p. 18210-9. Tu, X. and CC. Wang, Pairwise knockdowns of cdc2-related kinases (CRKs) in Trypanosoma brucei identified the CRKs for GI/S and GZ/M transitions and demonstrated distinctive cytokinetic regulations between two developmental stages of the organism. Eukaryot Cell, 2005. 4(4): p. 755-64. Vondruskova, E., et al., RNA interference analyses suggest a transcript-specific regulatory role for MRP1 and MRP2 in RNA editing and other RNA processing in trypanosoma brucei. J Biol Chem, 2004. Downey, N., et al., Mitochondrial DNA Ligases of Trypanosoma brucei. Eukaryot Cell, 2005. 4(4): p. 765-74. El-Sharoud, W.M. and G.W. Niven, The activity of ribosome modulation factor during growth of Escherichia coli under acidic conditions. Arch Microbiol, 2005. 184(1): p. 18-24. Ryan, C.M., K.T. Militello, and L.K. Read, Polyadenylation Regulates the Stability of Trypanosoma brucei Mitochondrial RNAs. The Journal of Biological Chemistry, 2003. 278(35): p. 32753-32762. 90 I"Ellililllljflljjillljjflll