<<

14 P

Sidney Altman Department of Molecular, Cellular and Developmental Biology Yale University New Haven, Connecticut 06520

Leif Kirsebom Department of Microbiology Biomedical Center Uppsala University Uppsala S-751 23, Sweden

RNase P was first characterized in Escherichia coli as an that is required for the processing of the 5 termini of tRNA in the pathway of biosynthesis of tRNA from precursor tRNAs (ptRNAs; Altman and Smith 1971; for review, see Altman 1989). Subsequently, it was discov- ered that this enzyme consisted of one and one RNA subunit, the latter being the catalytic subunit (Guerrier-Takada et al. 1983). In retro- spect, it should not be surprising that an enzyme with the chemical compo- sition of RNase P exists. are also ribonucleoproteins (RNPs), but they are much more complicated than RNase P. It seems unreasonable that ribosomes, with three and about 50 , would exist with- out less complex RNPs having come into existence first and having per- sisted throughout evolution. Indeed, relatively simple RNPs with and without catalytic activity have been discovered with regularity during the past 20 years. What important questions about RNase P have not yet been an- swered? Certainly, for the biochemist, problems not yet solved for E. coli RNase P include a full understanding of enzyme–substrate interac- tions, RNA–protein subunit interactions, and the chemical details of the hydrolysis reaction. In every case, much progress has been made, but much has yet to be learned. A second issue of interest to biochemists and “evolutionists” is the puzzle presented by the different compositions of RNase P from eubac- teria (one catalytic RNA and one protein subunit) and from (one RNA subunit and several protein subunits: no understanding yet of which subunit is responsible for catalysis). The complexity of eukaryotic

The RNA World, Second Edition © 1999 Cold Spring Harbor Laboratory Press 0-87969-561-7/99 351 352 S. Altman and L.A. Kirsebom

RNase P presents the questions of why there are so many subunits and what relation this complexity might have to the intracellular localization of the enzyme, possible isoforms (Li and Williams 1995), and its relation to other with which it might interact in vivo. RNase P in orga- nelles, e.g., and mitochondria, present their own complexities in terms of the origin and function of their subunits (Wang et al. 1988; Dang and Martin 1993; Baum et al. 1996; Lang et al. 1997; Martin and Lang 1997). More generally, the role of RNase P in the regulation of tRNA and rRNA biosynthesis in both prokaryotes and eukaryotes has not yet been fully explored. Indeed, new roles for RNase P in cell physiology are still being uncovered (Stolc and Altman 1997). This chapter focuses on enzyme–substrate interactions and summa- rizes briefly what is known about RNase P from (predominantly E. coli) that is pertinent to the other questions presented above and con- trasts the characteristics of E. coli RNase P with eukaryotic RNase P (pre- dominantly human RNase P). The discussion is not meant to be encyclo- pedic: We apologize to those whose work we have not mentioned.

FUNCTION OF RNASE P In E. coli, RNase P processes not only ptRNAs (Fig. 1) but also, at the very least, the precursors to 4.5S RNA (Bothwell et al. 1976a), an analog of the eukaryotic SRP RNA (Poritz et al. 1990), tmRNA (10Sa RNA; Komine et al. 1994; Keiler et al. 1996; Muto et al. 1996; Felden et al. 1997), and some small phage RNAs (Hartmann et al. 1995). Formally, this enzyme also has another substrate, 30S prRNA, which it cleaves at the 5 termini of tRNA sequences embedded in the long prRNA (for reviews, see Apirion and Miczak 1993; Deutscher 1993). There may very well be other “hidden” substrates, e.g., ones that resemble the poly- cistronic his operon mRNA, which have not yet been identified in the com- plexity of RNA processing events in vivo. In this latter case, a cleavage site by RNase P is accessible only after a previous cleavage of the mRNA by RNase E (Alifano et al. 1994). Only after the RNase E cleavage can the remaining 3 terminal fragment of the polycistronic mRNA adopt a con- formation that presents the recognition features of substrates for RNase P. To date, only one set of substrates for purified human RNase P has been identified: These are ptRNAs. However, in crude extracts of both and HeLa cells, RNase P appears to be part of a complex that includes RNase MRP and that is involved in the cleavage of prRNAs (Lygerou et al. 1996; Jarrous et al. 1998). Highly purified RNase P, how- ever, cannot carry out this latter cleavage reaction (Jarrous et al. 1998). Ribonuclease P 353

Figure 1 Natural substrates for RNase P from E. coli. Secondary structures of several substrates are shown. (A) Precursor to tRNATyr. (B) Precursor to 4.5S RNA. (C) Precursor to tmRNA. (D) Precursor to phages P1 and P7 antisense C4 RNA. (E) his operon mRNA at the junction of the his C and his B of Salmonella typhimurium (see text for explanation and references).

Identification of new substrates for eukaryotic RNase P should involve the examination of cell extracts depleted in RNase P activity. The fact that nuclear RNase P from both human and yeast cells have several protein subunits points to the possibility that these subunits, alone or together in the RNase P holoenzyme complex, have a function other than that of RNA cleavage. (RNase P from Saccharomyces cerevisiae and HeLa cells [Eder et al. 1997; Chamberlain et al. 1998] have been suffi- ciently highly purified to enable the previous statement regarding the number of subunits to be made. In S. cerevisiae, there is also genetic evi- dence for the identity of at least five subunits [for references, see Stolc and Altman 1997; Stolc et al. 1998].) In fact, the sequences of at least two of the human subunits have motifs resembling those found in proteins that are known to have nuclear localization signals (Jarrous et al. 1998). 354 S. Altman and L.A. Kirsebom

In human cells, RNase P must also be transported to the mitochondria (Doersen et al. 1985). (Unlike mitochondria in human cells, in several and other lower eukaryotes, the mitochondria appear to encode their own RNase P [Hollingsworth and Martin 1986; Martin and Lang 1997].) Accordingly, one of the protein subunits of the human enzyme may be important for the transport process. One candidate is the protein called p38 or Rpp38 (Eder et al. 1997), a nuclear antigen that is also found in the RNP enzyme known as RNase MRP (Yuan et al. 1991). As with RNase P, this enzyme is thought to have both nuclear (nucleolar) and mitochondrial function. The p38-binding domains in the RNA subunits of each of these enzymes are similar in structure (see Fig. 2) (Forster and Altman 1990b). Indeed, there are significant similarities in the overall proposed secondary structure of these RNAs. We note also that the two enzymatic functions are physically associated with each other in crude extracts of both human and yeast cells (Tollervey 1995; Stolc and Altman 1997), an indication of function in similar, or the same, biosynthetic path- ways (e.g., rRNA biosynthesis) and perhaps a common origin during evo- lution (see below). What are the common features of the various substrates that are rec- ognized by RNase P? Extensive genetic and biochemical analyses of both natural and model substrates indicate that the minimal features recognized by E. coli RNase P are a hydrogen-bonded stem of >3 bp and at least one unpaired nucleotide at the 5 end of the sequence (Liu and Altman 1996). The sequence –CCA at the 3 end of this stem (common to all mature tRNAs) enhances cleavage efficiency (Seidman and McClain 1975; Guerrier-Takada et al. 1984; Green and Vold 1988). The variety of known model substrates is shown in Figure 3. In contrast, human RNase P requires a somewhat more complex minimal structure for recognition by the enzyme (Fig. 3), one in which there is a bulge of one or more nucleotides at position 8 (Yuan and Altman 1995; see below).

SOME DETAILS OF THE REACTION WITH DIFFERENT SUBSTRATES The simplest way of studying an enzymatic reaction is to perform an analysis of its kinetics and to quantitate the well-known parameters, Km and kcat. The values of these parameters for several substrates for E. coli RNase P are listed in Table 1. Note that whereas some model sub- strates have a value of kcat/Km (a measure of the efficiency of the reaction) close to that of the enzyme with a natural substrate, the Km with ptRNA is the lowest, an indication that the interaction between ptRNA and the holoenzyme is stronger than with other substrates. We conclude, not Ribonuclease P

Figure 2 Secondary structure schemes of (A) M1 RNA and (B) H1 RNA. The of H1 RNA that is analogous to a domain to which the Th/To antigen, p38, binds (Yuan et al. 1991) in MRP RNA is shown in bold letters (Forster and Altman 1990b). (A, 355 Reprinted, with permission, from Chen and Pace 1997, B modified from Chen and Pace 1997.) 356 S. Altman and L.A. Kirsebom unexpectedly, that a natural substrate has more contacts with the enzyme than do model substrates. Indeed, mutagenesis studies show that any mutation that disturbs the secondary and tertiary structure of the tRNA moiety of the substrate in any part of the molecule alters the Km of the

Figure 3 Model substrates for E. coli (A–C) and human RNase P (D–F). (A) pAT1, (B) Target RNA-EGS (external guide sequence) complex. (C) Minimal substrate with base preferences shown. (D) Bulged hairpin (analog of pAT1) rec- ognized by human (Yuan and Altman 1995), X. laevis (Carrara et al. 1989), and Drosophila melanogaster (Levinger et al. 1997). (E) Target RNA-EGS complex (3/4 EGS). (F) Mini-EGS and target complex (Werner et al. 1997). Ribonuclease P 357

Table 1 Kinetic parameters of the reaction of M1 RNA and RNase P with various substrates

Km kcat kcat/Km Substrate Enzyme (nM) (min–1) (min–1/µM) pTyra M1 30 0.4 13 M1 + C5b 33 29 879 ∆65 0 ∆65 + C5 200 3 15 ∆[273-281] 97 1 10 ∆[94-204] 0 ∆[94-204] + C5 0 pTyr-CCAc M1 960 30 31 M1 + C5 110 101 920 ∆[273-281] 250 11 44 p4.5Sd M1 11,500 1 0.1 M1 + C5 150 60 400 ∆65 0 ∆65 + C5 0 ∆[273-281] 84,700 11 0.13 ∆[94-204] 0 ∆[94-204] + C5 1,170 44 37 p10AT1e M1 1,460 7 4.8 M1 + C5 1,010 32 32 ∆[273-281] 1,086 0.7 0.6 aThis natural substrate terminates in the sequence ACCAUCA. The presence of the extra three nucleotides has little effect on the kinetics of the cleavage reaction. bM1 RNA and C5 protein are in molar ratios of 1:10 in the reaction mixtures. cThis substrate terminates in A: The terminal CCAUCA sequence is missing. dThe precursor to 4.5S RNA of E. coli. eThe model substrate pAT1 (McClain et al 1987) with ten nucleotides in its 5 leader sequence. reaction (Kirsebom and Altman 1989; Svärd and Kirsebom 1992; Hardt et al. 1993; Kufel and Kirsebom 1994). Note also that the kcat of the reaction is enhanced by the presence of the protein , C5 protein. In certain cases, C5 protein also greatly reduces the Km, e.g., with p4.5S RNA (Peck-Miller and Altman 1991). C5 protein also enables the enzyme to discriminate among several tRNA substrates by differentially altering the relative rates of the cleavage reaction with different substrates (Kirsebom and Altman 1989; Gopalan et al. 1997). 358 S. Altman and L.A. Kirsebom

When ptRNAs have the common 3-terminal RCCA sequence, the rate-limiting step of the RNA alone reaction, as determined under multi- ple turnover conditions, is product release: Addition of the protein cofac- tor facilitates the release of the product (Reich et al. 1988; Tallsjö and Kirsebom 1993). Recently, a thorough analysis of the kinetics of the reac- tion with Bacillus subtilis RNase P RNA (a type B RNase P RNA; see Haas et al. 1996) has established a model that attempts to assign a kinetic parameter to every stage of the reaction, from substrate and ion binding to product release (Beebe and Fierke 1994). The model should be applicable, in most respects, to all bacterial RNase P RNAs, but some of its steps remain to be characterized more completely both in theory and in relation to experimental data. The 3-terminal sequence, –CCA, is found in mature tRNAs in all organisms. In eukaryotes, this sequence is absent from the tRNA tran- scripts but is added posttranscriptionally. It is present in all tRNA gene transcripts in E. coli and in almost all transcripts in other bacteria, too (Fleischmann et al. 1995; Fraser et al. 1995, 1997; Himmelreich et al. 1996; Blattner et al. 1997; Kunst et al. 1997; Tomb et al. 1997). The pres- ence of the CCA sequence in all ptRNAs in E. coli raises the question of whether this sequence plays a role in recognition of substrates by RNase P. (Note that p4.5S RNA ends in –CC and that the terminal A turns over in tRNAs and is not essential for recognition [Deutscher 1993 and references therein; see below].) The CCA sequence is important for recognition by M1 RNA but plays a lesser role when C5 protein is also present (see Tables 1 and 2). When the 5 C of the terminal two Cs is substituted by dC, the rate of the reaction is reduced about tenfold (Perreault and Altman 1992). The absence of the second C of the CCA sequence also decreases the rate of the reaction considerably (Table 2). Curiously, if the entire CCA sequence is missing, the rate of cleavage is not so severely affected, an indication, perhaps, that the other important features of the measuring and cleavage mechanisms are no longer affected by the identity or positioning of the 3-terminal sequence. These results have been derived from studies of both model and natural substrates from which terminal nucleotides have been individually and sequentially removed. Genetic studies have also determined that the CCA sequence hydrogen-bonds to a specific, comple- mentary sequence in M1 RNA (Kirsebom and Svärd 1994; Svärd et al. 1996; see below). This interaction also involves the discriminator base at position +73 in the ptRNA, since it has been shown that the absence of an interaction between the discriminator base and M1 RNA facilitates prod- uct release (Tallsjö et al. 1996). Accordingly, we can assign this sequence, Ribonuclease P 359

Table 2 The effect of changes in the 3-terminal CCA sequence on the kinetic parameters of the M1 RNA and RNase P cleavage reactions

Km kcat kcat/Km Substrate Enzyme (nM) (min–1) (min–1/µM) pTyra M1 33 0.4 12 M1 + C5b 29 76 260 pTyr-CAc M1 200 0.15 0.75 M1 + C5 125 18 140 pTyr-CCAd M1 500 3 6 M1 + C5 200 27 140 Yeast M1 100 0.44 4.4 pSupS1e M1 + C5 200 7.6 38

The experiments reported in Tables 1 and 2 were performed with different preparations of M1 RNA. aThis natural substrate terminates in the sequence ACCAUCA. The presence of the extra three nucleotides has little effect on the kinetics of the cleavage reactions. bM1 RNA and C5 protein are in molar ratios of 1:10 in the reaction mixtures. cTerminates in AC: The terminal CAUCA is missing. dTerminates in A: The terminal CCAUCA is missing. eDerived from S. cerevisiae tRNASer. This substrate does not have a 3-terminal CCA sequence. the unpaired nucleotide at the 5 side of the mature tRNA sequence (the “–1” nucleotide), and a short double-stranded region in the aminoacyl stem as part of the “minimal” recognition features (see Fig. 3C). Since RNase P requires a divalent metal ion, of which Mg++ is the most efficient as a cofactor (Gardiner et al. 1985; Guerrier-Takada et al. 1986; Kazakov and Altman 1991), there has been a great deal of interest in the number of divalent metal ions that are required for the hydrolysis reaction to proceed. The results of various studies suggest that two or three metal ions are required for catalysis (Kahle et al. 1993; Smith and Pace 1993; Warnecke et al. 1996; Chen et al. 1997). There are, apparently, several strong metal-ion-binding sites in M1 RNA but whether, and how many of, the ions bound to these sites are involved directly in the cleav- age event is not clear (Kazakov and Altman 1991; Zito et al. 1993; Ciesiolka et al. 1994). However, recent data suggest that a Mg++ ion is coordinated directly to the pro-Rp oxygen at the cleavage site and that this ion participates in the chemistry of cleavage (Warnecke et al. 1996; Chen et al. 1997). These recent data are consistent with another study which indicates that one of the required metal ions is bound to the substrate near 360 S. Altman and L.A. Kirsebom the site of cleavage and is carried to the active center of the enzyme in this manner (Perreault and Altman 1993). The substitution with chemical analogs of the usual nucleotides found in RNA has proved to be useful in determining which nucleotides are essential for catalysis in M1 RNA, its substrates, and in other . Many studies of E-S complexes have been performed in which each of these RNAs, randomly substituted with phosphorothioate backbones, deoxynucleotides or 4-thiouracil, are probed in E-S complexes with chemical agents that react with the RNAs (see below). Those phospho- rothioate bonds or substituted nucleotides that do not appear in the reac- tion products or that fail to react with the chemical agents are judged to be protected by E-S contacts. Experiments can also be carried out directly on unsubstituted RNAs. Finally, cross-linking agents can be attached to specific positions in “cyclized” M1 RNA or its substrate: Cross-links formed between RNAs must be in areas of close contact (~15 Å). The results of all these experiments, and others carried out with deoxy-substituted model substrates, can be summarized as follows:

1. The aminoacyl stem, or its equivalent in a substrate, is denatured dur- ing catalysis by M1 RNA, probably after binding of the substrate but before hydrolysis occurs. These data are derived from protection from chemical reagents of a ptRNA substrate in E-S complexes (Knap et al. 1990). 2. Deoxynucleotide substitutions at various positions in model and ptRNA substrates have been used to show that positions –1 and –2, as well as the 5-terminal C of the CCA sequence (Perreault and Altman 1992; Kleineidam et al. 1993; Kufel and Kirsebom 1996a), severely affect the kcat of hydrolysis. Similar substitutions in other parts of tRNA substrates show that binding to M1 RNA is critically affected by the loss of the 2-OH at G + 1, and in the T stem and T loop (Gaur and Krupp 1993; Conrad et al. 1995; Loria and Pan 1997). 3. Phosphorothioate substitutions have been used to show that the aminoacyl and T stems of substrates unfold during the reaction, as well as to demonstrate that ptRNAs with either a short or a long vari- able arm interact differently with M1 RNA (Gaur et al. 1996). This approach has also been used to identify phosphates and 2-OH in RNase P RNA important for tRNA binding (Hardt et al. 1995, 1996) and catalysis (Harris and Pace 1995). Here, 2-OH and phosphates important for tRNA binding have been localized at and near the P4 helix as well as to residue C247 in M1 RNA (see Fig. 2). Phosphates in P4 are also thought to be important for Mg++ binding and catalysis. Ribonuclease P 361

4. Chemical cross-linking agents have been used in various ways to show that the 3-terminal CCA sequence, the aminoacyl stem, the T stem (positions 49–53 and 61–65), and loop are critical for binding of a tRNA precursor to M1 RNA (Nolan et al. 1993; Harris et al. 1994; Oh and Pace 1994) as well as to identify nucleotides in M1 RNA that are in close contact with its substrate (see below; Guerrier-Takada et al. 1989; Burgin and Pace 1990; Harris et al. 1994; Kufel and Kirsebom 1996a; Massire et al. 1998). However, some of these experiments have been carried out with tRNAs rather than ptRNAs. It is not yet clear that both ptRNAs and tRNAs bind in precisely the same manner to M1 RNA. Other data indicate that product inhibition of the reaction is noncompetitive (Guerrier-Takada and Altman 1993), but there are some differences in experiments of this kind from different laboratories. Through the use of various substrates, differ- ences in the way in which M1 RNA interacts with particular sub- strates have been observed both by cross-linking and chemical pro- tection studies (Guerrier-Takada et al. 1989; Kufel and Kirsebom 1996a,b; Pan and Jakacka 1996; see also section 3 above). Neverthe- less, many of the data agree with results cited in sections 2 and 3 above and are compatible, to a large extent, with three-dimensional models of M1 RNA–ptRNA interactions.

Human RNase P, as mentioned above, has more constraints in recog- nizing its substrates than does E. coli RNase P, with one exception: The 3-terminal–CCA sequence is unimportant in determining enzyme effi- ciency. We note also that eukaryotic RNase P RNAs lack the region in the internal loop of P15 in the secondary structure scheme depicted in Figure 2. These RNAs, as well as their analogs from Chlamydia spp. and some cyanobacteria, do not need the specific hydrogen-bonded interaction with the –CCA sequence found in E. coli M1 RNA (Herrmann et al. 1996; Vioque 1997). Hardt et al. (1993) have also shown that, unlike hairpin substrate for RNase P from E. coli, RNase P from HeLa cells requires some tertiary structure at the junction between the equivalents of the aminoacyl and the T stems in its substrates for recognition to occur. These results were confirmed by Yuan and Altman (1995) and Werner et al. (1997). Both substrate affinity and efficient catalysis by S. cerevisiae RNase P appear to be influenced by the presence of a mismatch of the nucleotide at –1 (Lee et al. 1997). Furthermore, in the absence of the 3-terminal CCA sequence, if the nucleotide at –1 is base-paired with a nucleotide at the 3 end of the ptRNA, denaturation of the aminoacyl stem to facilitate 362 S. Altman and L.A. Kirsebom positioning and cleavage by RNase P becomes energetically more diffi- cult (see below). Since no eukaryotic RNase P RNA has been shown to be catalytic in vitro, it is likely that the protein subunits of these RNase Ps play a greater role in substrate recognition than does the protein (subunit) of E. coli RNase P.

SEQUENCE AND CLEAVAGE-SITE RECOGNITION IN SUBSTRATES RNase P does not recognize specific sequences in its substrates. From studies of RNase P from E. coli, we conclude that there are, however, some general preferences of the enzyme for certain nucleotides in the minimal recognition domain. Not surprisingly, these preferences reflect the naturally occurring composition of tRNAs. For example, the vast majority of mature tRNAs have a G in the first position of their sequences, a high G-C content in the aminoacyl stem, U in position 8 and, as discused above, a 3-terminal CCA sequence. Some data suggest that G at –2 also influences both cleavage efficiency and cleavage-site recognition (Kirsebom and Svärd 1993; Meinnel and Blanquet 1995; M. Bränvall et al., unpubl.). The efficiency of the RNase P cleavage reaction in ptRNAs is frequently decreased considerably if any of these features of substrates are altered. Another interesting aspect of the reaction, namely miscleav- age, is revealed when G1 is changed, or the base-pairing near the ends of the aminoacyl stem is destroyed (see, e.g., Kirsebom 1995 and references therein). One of the early questions posed in studies of RNase P concerns the accuracy of the cleavage reaction; i.e., how does the enzyme always find the right cleavage site in a large number of substrates that differ in sequence? Put another way, we can ask if RNase P has a measuring device that senses the tRNA domain of ptRNAs and, accordingly, cleaves a cer- tain distance from particular features in this domain. These features include the structural integrity of the T loop, the length of the T stem and the length of the aminoacyl stem, commonly 7 bp (with two exceptions; see below). The enzyme appears to measure the distance, in almost every case 12 bp between the T loop and the cleavage site (Kahle et al. 1990; Thurlow et al. 1991; Svärd and Kirsebom 1992, 1993; Kufel and Kirsebom 1994, 1996b). Additionally, the RCCA at the 3 termini of ptRNAs has a specific role in this process (see below). The physical attributes of these various features of M1 RNA have to operate in concert in order to accomplish cleavage of the substrate at the correct position. We Ribonuclease P 363 propose, specifically, that the enzyme interacts with the T loop (P11 and A118 in M1 RNA; G64 in the substrate with A118 and C100 in M1 RNA; see Figs. 4 and 5) as proposed by Massire et al. (1998) a certain distance in the substrate, equivalent to 12 bp, from where the actual cleavage event takes place. This has to occur in concert with the RCCA–RNase P RNA interaction (with the conserved GGU in the internal loop of P15; interact- ing residues as shown in Figure 6 as suggested by Kirsebom and Svärd 1994; see also LeGrandeur et al. 1994; Oh and Pace 1994; Svärd et al. 1996; Tallsjö et al. 1996). A third contact point between the RNase P RNA (A249; see Fig. 4) and its substrate would be an interaction with, in particular, a G at the cleavage site (+1) and perhaps also with the nucleotides at positions –1 and –2. The outcome of these interactions would be to expose the cleavage site and consequently correctly position the scissile bond in the (A351, A352, J18/2, P4; see Fig. 4). This model can also account for the correct cleavage of various model substrates, since these substrates still exhibit at least two of the features important for cleavage-site recognition. RNase P RNA folding would generate a pocket that can accommodate the T loop and the coaxially stacked T stem and aminoacyl stem as shown in Figure 4. The stem region, P5, is a likely candidate for part of the “measuring” apparatus. The exact contacts between substrates and RNase P RNA from dif- ferent species might change according to the nature of the RNA from that species. We would expect that RNase P RNA and its substrates would have co-evolved to preserve the necessary interactions (see below). (Sequence analysis reveals that the tRNA genes in Borrelia burgdorferei [Fleischmann et al. 1995] do not encode CCA despite the presence of the P15 loop in its RNase P RNA. This may be an example of a branch point in the evolution of RNase P RNA.) Evidence supporting our picture of a measuring mechanism comes from studies with model substrates in which extra base pairs have been added to or deleted from the aminoacyl and/or the T stem in combination, in some cases with removal of the terminal RCCA sequence (Svärd and Kirsebom 1992, 1993; Kufel and Kirsebom 1994, 1996b). Addition of 1 or 2 bp in the analogs of the aminoacyl and T stems has no effect on pre- cision of cleavage (Svärd and Kirsebom 1992). In these cases, the preci- sion of cleavage depends, we assume, mainly on the RCCA–RNase P RNA interaction as well as the nucleotide at the cleavage site. When 3 bp are inserted (Svärd and Kirsebom 1993), cleavage occurs at the 5 side of positions +1 (the 5 side of the first base pair in the altered aminoacyl stem) and +4, an indication that both aspects of the measuring device are 364 S. Altman and L.A. Kirsebom

Figure 4 Computer-assisted modeling of M1 RNA (Massire et al. 1998). Pro- posed secondary structure (A) and three-dimensional ribbon model of M1 RNA (B). Color coding of domains is the same for each domain shown in (A) and in (B). (Reprinted, with permission from Massire et al. 1998 [copyright Academic Press].) (C, D) Two views of a three-dimensional ribbon model of M1 RNA (white) and bound precursor tRNAAsp (red). Positions +1 (cleavage site), 64 and 53 (T stem), from bottom to top, in the tRNA domain are fixed accurately (cyan blue) as refer- ence points for the interaction with the enzyme (M1 RNA, in this case). Ribonuclease P 365

Figure 5 Scheme for the “measuring” device in, and selection of cleavage site in a tRNA precursor by, RNase P RNAs. (A) Bacterial RNase P RNA. (B) Eukaryotic RNase P RNA. The common features of each scheme are the domains that interact with the T loop, that “measure” the distance between the T loop and the cleavage site, that encompass the active center of the enzyme, and that inter- act with nucleotides in the vicinity of the cleavage site (+1 to –2). The active cen- ter also includes part of the substrate, since a Mg++ ion is coordinated to the Rp oxygen at the site of cleavage (Perreault and Altman 1993; Warnecke et al. 1996; Chen et al. 1997). The RCCA sequence in substrates interacts with a P15 domain found only in bacterial RNase P RNAs and, in particular, in those bacteria in which the RCCA sequence is encoded in the genomic DNA of tRNA genes. Additional data on which this model is based are given in the text. Bothwell et al. (1976b) offered a much simpler version of this scheme. 366 S. Altman and L.A. Kirsebom

Figure 6 P15-RCCA interactions between M1 RNA and precursor tRNA. (A) R in the substrate denotes an A or a G and results in an A-U or a G-U base pair in the enzyme–substrate complex. The figure also shows three Mg++ ions (yellow), one carried in by the substrate and two bound to the P15 loop in M1 RNA. (B) Structure of a model RNA that contains the P15 loop (Glemarec et al. 1996). The arrows denote the highly conserved GGU-motif in RNase P RNAs that base-pairs with the 3-terminal RCCA sequence of a tRNA precursor. Ribonuclease P 367 operative. If the 3-terminal RCCA is deleted, cleavage occurs almost exclusively at the +4 site (Kufel and Kirsebom 1994). Thus, there is some play in the correlation between the length of the aminoacyl and T stems and the point at which the measuring device stops relying solely on the RCCA and T-loop reference points for positioning of substrates. (We note that p4.5S RNA, which has a very long stem, is lacking a T-loop equiva- lent. Thus, the precision of cleavage must depend primarily on recogni- tion features near the site of cleavage on the 5 strand and the 3-terminal CC sequence.) In contrast to the situation with E. coli RNase P, changing the num- ber of base pairs in substrates for human or Xenopus Laevis RNase P by even 1 bp does affect cleavage precision in the sense that the enzyme always cleaves at a position 7 bp from the junction with the T stem (Carrara et al. 1989; Yuan and Altman 1995). It appears that the bacterial enzyme has sufficient information in a model substrate to determine where G1 is, but the human enzyme needs additional contacts in the T stem-loop to make this determination. It follows that this phenomenon might be explained in part by the absence of the RCCA–RNase P RNA interaction in eukaryotes. It is, indeed, possible to make RNase P alter the site of cleavage in bacterial substrates other than by adding 3 bp to the combined length of the aminoacyl and T stems (see below). The substitution of only one nucleotide within one of the minimal substrate recognition domains in certain substrates can also achieve this effect (Kirsebom and Svärd 1994; see below). The introduction of these structural features into a substrate overrides the measuring mechanism of even the minimal substrate recog- nition features. Changes of G1, or base-pairing at the end of the aminoacyl stem, can, in some cases, affect the choice of cleavage site. This phenomenon is dra- matically illustrated with E. coli ptRNAHis. tRNAHis is the one of two tRNAs that has 8 bp in its aminoacyl stem: the other is tRNASeCys. In bac- teria, the 8-bp stem in these tRNAs results from RNase P cleavage at a position that would normally be position –1 in other ptRNAs. Replacement of the guanosine at the ptRNAHis cleavage site or substitu- tion of C73 at the 3 end results in miscleavage by M1 RNA alone or by reconstituted holoenzyme such that 5 matured tRNAs with 7 bp forming an aminoacyl stem are produced. However, addition of C5 protein to the reaction mixture results in an increased frequency of cleavage at the cor- rect position to yield an aminoacyl stem of 8 bp (Burkard et al. 1988; Green and Vold 1988; Kirsebom and Svärd 1992). This is consistent with 368 S. Altman and L.A. Kirsebom the fact that details of the 5 and 3 termini of the tRNA domain affect the cleavage-site recognition and the measuring process. In contrast, eukary- otic RNase P cleavage of ptRNAHis generates a 7-bp aminoacyl stem and the extra G is added after RNase P cleavage (Cooley et al. 1982). This phenomenon is again in keeping with our speculations concerning mea- suring and the absence of the CCA at the 3 termini of eukaryotic tRNA precursors, as well as the lack of a P15 loop equivalent, the site of RCCA interaction, in eukaryotic RNase P RNA.

3-D MODELS AND SUBSTRATE RECOGNITION Any successful model of RNase P must be able to account for the data gathered so far on substrate binding and the detailed mechanism of the cleavage reaction. The latter, unfortunately, is still largely unknown. In the absence of a crystal structure of the enzyme, theoretical, computer- assisted models (Harris et al. 1994, 1997; Westhof and Altman 1994) derived from phylogenetic and experimental data have been useful in leading the way to further experimentation. The most recent model (Fig. 4) (Massire et al. 1998) is derived primarily from phylogenetic analysis but is also consistent with a large fraction, but not all, of the data available from chemical, cross-linking studies and genetic studies. It is more com- patible with Fe-EDTA protection data for M1 RNA (Westhof et al. 1996) than are previous models (D. Wesolowski and S. Altman, unpubl.). A compact, catalytic center for M1 RNA is a feature of the new model. In its organization of RNA domains, it shares some features with the catalytic center of group I introns. Nevertheless, there are still some discrepancies between the proposal in the model that certain structures (e.g., that are encompassed by nucleotides 94–204) are essential for catal- ysis and experimental data which show they can be deleted from M1 RNA molecules that still retains function in vitro. Although the precise details of the orientation of a ptRNA substrate on the surface of M1 RNA remain to be worked out, it is clear that the model generally agrees with what is known about contacts with the T stem and loop and the RCCA domain of substrates (see above). These constraints, themselves, give credence to the model and also allow it to be considered as a platform on which to dock and test experimental data regarding the conformation and binding to M1 RNA of the C5 protein cofactor (Gopalan et al. 1998).

NATURAL HISTORY OF RNASE P Ribonuclease P 369

Several groups have shown that M1 RNA can be linked to other RNAs and still retain its catalytic function. Three kinds of constructs have been made: (1) M1 RNA covalently linked at either terminus to a ptRNA or model substrate (Kikuchi et al. 1993; Frank et al. 1994; Kikuchi and Suzuki-Fujita 1995); (2) M1 RNA linked at either terminus to an external guide sequence (EGS: i.e., the 3-terminal segment of the aminoacyl stem of a tRNA [Li et al. 1992; Yuan et al. 1992; Guerrier-Takada et al. 1995, 1997; Liu and Altman 1995, 1996; Li and Altman 1996]); and (3) M1 RNA covalently linked to a ptRNA followed by a portion of the bacterio- phage T4 gene, thymidylate synthase (td), which contains two exons sep- arated by an intron (A. Kaplin et al., unpubl. cited in Altman 1989). In these cases, individually, cleavage was observed at the correct site in (1) the ptRNA segment; (2) the target sequence of an EGS-target complex that resembles part of a ptRNA (i.e., the 5 segment of a ptRNA amino- acyl stem and the upstream leader sequence; and (3) in the ptRNA sequence concomitantly with self-splicing of the partial td gene. We note that the M1 RNA-ptRNA or M1 RNA-EGS RNA construct reactions are quite efficient. Presumably, binding of the potential substrate to the active center of M1 RNA is facilitated by the tether linking it to the enzyme (the tether must be longer than a minimum length, about 7 nucleotides [Liu and Altman 1996]). These experiments suggest that complex RNAs can exist, and could have existed eons ago in an RNA World, which contain two or more independent catalytic activities and substrates (see Fig. 7A). Contemporary M1 RNA and its substrates must have been somewhat simpler at one point in time than they are today. The reasoning for this statement is presented below, but there is an implicit assumption that should be noted first: We equate “simpler” with “earlier” in time of evo- lution. Such an assumption depends in a not very well-defined manner on the meaning of the word “simpler.” Working with purely phylogenetic analyses, Pace and coworkers have defined what they call the “minimal” RNase P RNA, i.e., the smallest molecule (211 nucleotides; Siegel et al. 1996) needed to carry out the hydrolysis reaction associated with RNase P. Furthermore, an even smaller RNase P RNA (140 nucleotides) has been identified in the mitochondria of Saccharomyces fibuligera (Wise and Martin 1991), although this RNA is not catalytic by itself. One can also demonstrate in another fashion that M1 RNA must have been smaller at a point in time earlier than today. Contemporary M1 RNA has two basic classes of substrates, exemplified by precursor tRNA and precursor 4.5S RNA, respectively. The latter sub- strate is structurally the simpler of the two, since it contains only a hair- pin and an upstream leader sequence. The Km of the reaction of M1 RNA 370 S. Altman and L.A. Kirsebom

(not including C5 protein) with each substrate is an indication of the com- plexity of each substrate: The Km with pTyr is about 100-fold smaller than with p4.5S RNA, a reflection of the larger number of contacts made with the former substrate. Furthermore, the entire “upper” domain of M1 RNA (nucleotides 94–204; see Fig. 2) can be deleted without a major effect on

Figure 7 Scheme for the early evolution of RNA and RNPs. (A) Long RNA molecule with occasional hairpins and a sequence that contains an ancient M1 RNA-like (ur-M1) cleavage activity. (B) Molecule in A is self-cleaved at 5 side of hairpins by ur-M1. (C) Beginning of the progression of coevolution of M1 RNA, substrates, and protein cofactors. (D1, D2) Branch points leading to eukary- otic RNase P and RNase MRP. Ribonuclease P 371 the reaction with p4.5S RNA, whereas the reaction with various ptRNAs is less efficient when this domain is lost. We conclude, therefore, that a simpler form of today’s M1 RNA, dictated by either Pace’s criterion or the one outlined immediately above, must have existed in which the upper domain was absent, since it is not needed for the binding of “hairpin” sub- strates (the simpler class of substrates) or catalysis. We recall that the sim- plest model substrates for M1 RNA consist only of a small hairpin and a few extra nucleotides at the 5 terminus. It is not irrelevant to mention that the simplest substrates for tRNA synthases are also small hairpins (Francklyn et al. 1992). In both cases, i.e., aminoacylation and RNase P cleavage, the Km values for the reactions with the model substrates are very high compared to reactions with today’s more complex, natural substrates. We speculate, accordingly, that the original role of an ancient M1 RNA may have been to cleave long RNAs at the 5 side of hairpin structures to generate a variety of smaller molecules (Fig. 7A, B). (Maizels and Weiner [1993] also recognized a possible important early role of hairpin structures in their development of the “genome tagging” theory of RNA replication.) The population of more numerous smaller molecules should be endowed with a larger num- ber of functional capabilities than might be possible with the original, par- ent, larger RNA molecule. One can imagine that greater functional varia- tion comes in part from discrete structural domains of RNA interacting with each other, not necessarily through covalent bonds, to create larger complexes with active functions and to present even more possibilities for assembly of an ancient M1 RNA with cleavage capability. Indeed, com- plexes with M1 RNA activity have been constructed in vitro from sub- domains with no catalytic activity (Guerrier-Takada and Altman 1992; Pan 1995). We do not assign an absolute time to the origins of M1 RNA and its substrates. The possibility that the first self-replicating genetic system was not made of RNA, or the specific place of M1 RNA in an evolutionary time series in the RNA World is not particularly relevant. The only impor- tant assertion we make in a discussion of the evolution of M1 RNA and other catalytic RNAs and their attendant proteins is that there was an RNA World at some time prior to the appearance of proteins as entities encoded in genetic material. Furthermore, given the uncertainties in “replication times” of the inhabitants of the RNA World, the nature of the then-contemporary selective pressures and nucleotide substitution rates, it seems of little value to attempt to assign rough, absolute time intervals to different stages of complexity of the RNA World. Notwithstanding the uncertainties cited above, it is still possible to 372 S. Altman and L.A. Kirsebom consider the nature and fate of individual molecules in, and their progres- sion during, evolution to a resemblance of those that we see today. For example, an evolutionary scheme can be described in which we consider a set of RNA–protein complexes that represent a progression in an evolu- tionary sense, and that might contain an ancient M1-like catalytic RNA (a “complex” with one RNA and no protein, R1P0), M1-like catalytic RNA plus a protein cofactor (R1P1), and ultimately, 16S rRNA plus 20 proteins (R2P2–21) if 16S rRNA is a direct evolutionary descendant of M1 RNA and its proteins are the immediate descendants of P1. We note that this particular progression is unlikely in reality: It is apparent that there must be more intermediates in the series, as noted above, in the evolution of M1 RNA itself (i.e., R1P0 . . . RNP0; see Fig. 7) as well as branch points in which the original R1 or its predecessor evolved into another molecule, R2, that might have been the evolutionary precursor of the self-splicing intron (or possibly, R3, a molecule that generated different end groups during cleavage and was the precursor of the class of catalytic RNAs known as viroids). Each of these three progenitor RNA molecules must have evolved further in terms of catalytic specificity and, ultimately, in their ability to work together with one or more proteins (see Fig. 7). In addition to these catalytic RNAs, a parallel evolution of noncat- alytic RNAs, e.g., tRNA, tmRNA, rRNA, is required, and these RNAs might (since some were substrates for the catalytic RNAs) or might not be co-varying with the catalytic RNAs. In fact, today we can see an example of coevolution of an RNA enzyme and its substrates in the ability of RNase P RNAs from two different sources, the so-called type A or type B RNase P RNAs, to cleave or not to cleave, respectively, p4.5S RNA from E. coli (Guerrier-Takada et al. 1983). Once proteins were on the scene, coevolution of catalytic RNAs, pro- teins, and substrate RNAs (Liu and Altman 1994) had to have been the general case. A clearer example of evolutionary progression is, perhaps, more obvious if we consider the lineage of RNase P from an early pro- genitor organism to E. coli RNase P to human RNase P with a relatively early branch point (perhaps at the same juncture in time as the branch between eubacteria and eukaryotes) that leads to RNase MRP (Chang and Clayton 1987), an enzyme found only in eukaryotes but which has a close relationship to RNase P, as noted previously (Fig. 7C, D). The thread of evolution can also be discerned in the observation that M1 RNA functions in mammalian cells in tissue culture, possibly by co-opting a protein sub- unit of host cell RNase P (Liu and Altman 1995). Is it possible to make a statement concerning the “time” at which pro- teins began to interact with RNA and infer something about the state of Ribonuclease P 373 the RNA World when proteins first appeared? We use as examples the facts that tRNA synthases can aminoacylate micro-hairpin substrates (Francklyn et al. 1992), and that RNase P can cleave similar, model ptRNAs (McClain et al. 1987; Forster and Altman 1990a). If these exam- ples bear any relation to today’s natural world, then in the cases of tRNA synthases and of RNase P, and very likely of other RNPs as well, RNA evolution, catalytic and otherwise, was far from finished when proteins arrived on the scene. One prediction from these ideas is that fragments of tRNA synthases and C5 protein will be able to carry out, or participate in as a cofactor, respectively, enzymatic reactions on micro-helical, model substrates.

ACKNOWLEDGMENTS We thank our colleagues for comments on the manuscript, and we thank D. Wesolowski and D. Pomeranz-Krummel for help with the figures. Work in the laboratory of S.A. is funded by U.S. Public Health Service grant GM-19422 and a Human Frontiers of Science program grant RG- 0291. Work in the laboratory of L.K. is supported by the Swedish Natural Science Research Council.

REFERENCES Alifano P., Rivellini F., Piscitelli C., Arraiano C.M., Bruni C.B., and Carlomagno M.S. 1994. provides substrates for ribonuclease P-dependent processing of a polycistronic mRNA. Genes Dev. 8: 3021–3031. Altman S. 1989. Ribonuclease P: An enzyme with a catalytic RNA subunit. Adv. Enzymol. Relat. Areas Mol. Biol. 62: 1–36. Altman S. and Smith J.D. 1971. Tyrosine tRNA precursor molecule polynucleotide sequence. Nat. New Biol. 233: 35–39. Apirion D. and Miczak A. 1993. RNA processing in prokaryotic cells. BioEssays 15:113–120. Baum M., Cordier A., and Schön A. 1996. RNase P from a photosynthetic organelle con- tains an RNA homologous to the cyanobacterial counterpart. J. Mol. Biol. 257: 43–52. Beebe J.A. and Fierke C.A. 1994. A kinetic mechanism for cleavage of precursor tRNAAsp catalysed by the RNA component of Bacillus subtilis ribonuclease P. Biochemistry 33: 10294–10304. Blattner F.R., Plunkett G., III, Bloch C.A., Perna N.T., Burland V., Riley M., Collado- Vides J., Glasner J.D., Rode C.K., Mayhew G.F., Gregor J., Davis N.W., Kirkpatrick H.A., Goeden M.A., Rose D.J., Mau B., and Shao Y. 1997. The complete genome sequence of Escherichia coli K-12. Science 277: 1453–1462. Bothwell A.L.M., Garber R.L., and Altman S. 1976a. Nucleotide sequence and in vitro processing of a precursor molecule to Escherichia coli 4.5 S RNA. J. Biol. Chem. 251: 374 S. Altman and L.A. Kirsebom

7709–7716. Bothwell A.L.M., Stark B.C., and Altman S. 1976b. Ribonuclease P substrate specificity: Cleavage of a bacteriophage f80-induced RNA. Proc. Natl. Acad. Sci. 73: 1912–1916. Burgin A.B. and Pace N.R. 1990. Mapping the active site of ribonuclease P RNA using a substrate containing a photoaffinity agent. EMBO J. 9: 4111–4118. Burkard U., Willis I., and Söll D. 1988. Processing of histidine transfer RNA precursors. J. Biol. Chem. 263: 2447–2451. Carrara G., Calandra P., Fruscoloni P., Doria M., and Tocchini-Valentini G. 1989. Site selection by Xenopus laevis RNase P. Cell 58: 37–45. Chamberlain J.R., Lee Y., Lane W.S., and Engelke D.R. 1998. Purification and character- ization of the nuclear RNase P holoenzyme complex reveals extensive subunit overlap with RNase MRP. Genes Dev. 12: 1678–1690. Chang D. and Clayton D.A. 1987. A novel cleaves at a priming site of mouse mitochondrial DNA replication. EMBO J. 6: 409–417. Chen J.-L. and Pace N.R. 1997. Identification of the universally conserved core of ribonu- clease P RNAs. RNA 3: 557–560. Chen Y., Li X., and Gegenheimer P. 1997. Ribonuclease P catalysis requires Mg2+ coor- dinated to the pro-Rp oxygen of the scissile bond. Biochemistry 36: 2425–2438. Ciesiolka J., Hardt W.-D., Schlegl J., Erdmann V.A., and Hartmann R.K. 1994. Lead-ion- induced cleavage of RNase P RNA. Eur. J. Biochem. 219: 49–56. Conrad F., Hanne A., Gaur R.K., and Krupp G. 1995. Enzymatic synthesis of 2-modified nucleic acids: Identification of important phosphate and ribose moieties in RNase P substrates. Nucleic Acids Res. 23: 1845–1853. Cooley L., Appel B., and Söll D. 1982. Post-transcriptional nucleotide addition is respon- sible for the formation of the 5 terminus of histidine tRNA. Proc. Natl. Acad. Sci. 79: 6475–6479. Dang Y.L. and Martin N.C. 1993. Yeast mitochondrial RNase P: Sequence of the RPM2 gene and demonstration that its product is a protein subunit of the enzyme. J. Biol. Chem. 268: 19791–19796. Deutscher M.P. 1993. RNA maturation . In Nucleases, 2nd edition (ed. S.M. Linn et al.), pp. 377–406. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, New York. Doersen C.J., Guerrier-Takada C., Altman S., and Attardi G. 1985. Characterization of an RNase P activity from HeLa cell mitochondria. Comparison with the cytosol RNase P activity. J. Biol. Chem. 260: 5942–5949. Eder P.S., Kekuda R., Stolc, V., and Altman S. 1997. Characterization of two scleroderma autoimmune antigens that copurify with human ribonuclease P. Proc. Natl. Acad. Sci. 94: 1101–1106. Felden B., Himeno H., Muto A., McCutcheon J.P., Atkins J.F., and Gesteland R.F. 1997. Probing the structure of the E. coli 10Sa RNA (tmRNA). RNA 3: 89–103. Fleischmann R.D., Adams M.D., White O., Clayton R.A., Kirkness E.F., Kerlavage A.R., Bult C.J., Tomb J.F., Dougherty B.A., Merrick J.M., et al. 1995. Whole-genome random sequencing and assembly of Haemophilus influenzae Rd. Science 269: 496–512. Forster A.C. and Altman S. 1990a. External guide sequences for an RNA enzyme. Science 249: 783–786. ———. 1990b. Similar cage-shaped structures for the RNA components of all ribonucle- Ribonuclease P 375

ase P and ribonuclease MRP enzymes. Cell 62: 407–409. Francklyn C., Shi J.P., and Schimmel P. 1992. Overlapping nucleotide determinants for specific aminoacylation of RNA microhelices. Science 255: 1121–1125. Frank D.N., Harris M.E., and Pace N.R. 1994. Rational design of self-cleaving pre-tRNA- ribonuclease P RNA conjugates. Biochemistry 33: 10800–10808. Fraser C.M., Casjens S., Huang W.M., Sutton G.G., Clayton R., Lathigra R., White O., Ketchum K.A., Dodson R., Hickey E.K., Gwinn M., Dougherty B., Tomb J.F., Fleischmann R.D., Richardson D., Peterson J., Kerlavage A.R., Quackenbush J., Salzberg S., Hanson M., van Vugt R., Palmer N., Adams M.D., Gocayne J., Venter J.C., et al. 1997. Genomic sequence of a Lyme disease spirochaete, Borrelia burgdor- feri. Nature 390: 580–586. Fraser C.M., Gocayne J.D., White O., Adams M.D., Clayton R.A., Fleischmann R.D., Bult C.J., Kerlavage A.R., Sutton G., Kelley J.M., Fritchman J.L., Weidman J.F., Small K.V., Sandusky M., Fuhrmann J., Nguyen D., Utterback T.R., Saudek D.M., Phillips C.A., Merrick J.M., Tomb J.-F., Dougherty B.A., Bott K.F., Hu P.-C., and Lucier T.S. 1995. The minimal gene complement of Mycoplasma genitalium. Science 270: 397–403. Gardiner K.J., Marsh T.L., and Pace N.R. 1985. Ion dependence of the Bacillus subtilis RNase P reaction. J. Biol. Chem. 260: 5415–5419. Gaur R.K. and Krupp G. 1993. Modification interference approach to detect ribose moi- eties important for the optimal activity of a . Nucleic Acids Res. 21: 21–26. Gaur R.K., Hahne A., Conrad F., Kahle D., and Krupp G. 1996. Differences in the inter- action of Escherichia coli RNase P RNA with tRNA containing a short or a long extra arm. RNA 2: 674–681. Glemarec C., Kufel J., Foldesi A., Sandstrom A., Kirsebom L.A., and Chattopadhyaya J. 1996. The NMR structure of 31 mer RNA domain of Escherichia coli RNase P RNA using its non-uniformly deuterium labelled counterpart [the ‘NMR-window’ concept]. Nucleic Acids Res. 24: 2022–2035. Gopalan V., Baxevanis A., Landsman D., and Altman S. 1997. Analysis of the functional role of conserved residues in the protein subunit of ribonuclease P from Escherichia coli. J. Mol. Biol. 267: 818–829. Gopalan V., Ledman D.W., Fox R.O., and Altman S. 1998. Mapping RNA-protein inter- actions in ribonuclease P from Escherichia coli using disulfide linked EDTA-Fe. J. Mol. Biol. (in press). Green C.J. and Vold B.S. 1988. Structural requirements for processing of synthetic tRNAHis precursors by the catalytic RNA components of RNase P. J. Biol. Chem. 263: 652–657. Guerrier-Takada C. and Altman S. 1992. Reconstitution of enzymatic activity from frag- ments of M1 RNA. Proc. Natl. Acad. Sci. 89: 1266–1270. ———. 1993. A physical assay for and kinetic analysis of the interactions between M1 RNA and tRNA precursor substrates. Biochemistry 32: 7152–7161. Guerrier-Takada C., Li Y., and Altman S. 1995. Artificial regulation of in Escherichia coli by RNase P. Proc. Natl. Acad. Sci. 92: 11115–11119. Guerrier-Takada C., Lumelsky N., and Altman S. 1989. Specific interactions in RNA enzyme-substrate complexes. Science 286: 1578–1584. Guerrier-Takada C., McClain W.H., and Altman S. 1984. Cleavage of tRNA precursors by the RNA subunit of E. coli ribonuclease P (M1 RNA) is influenced by 3-proximal CCA in the substrates. Cell 38: 219–224. 376 S. Altman and L.A. Kirsebom

Guerrier-Takada C., Salavati R., and Altman S. 1997. Phenotypic conversion of drug- resistant bacteria to drug sensitivity. Proc. Natl. Acad. Sci. 94: 8468–8472. Guerrier-Takada C., Haydock K., Allen L., and Altman S. 1986. Metal ion requirements and other aspects of the reaction catalyzed by M1 RNA, the RNA subunit of ribonu- clease P from Escherichia coli. Biochemistry 25: 1509–1515. Guerrier-Takada C., Gardiner K., Marsh T., Pace N., and Altman S. 1983. The RNA moi- ety of ribonuclease P is the catalytic subunit of the enzyme. Cell 35: 849–857. Haas E.S., Banta A.B. Harris J.K., Pace N.R., and Brown J.W. 1996. Structure and evolu- tion of ribonuclease P RNA in gram-positive bacteria. Nucleic Acids Res. 24: 4775–4782. Hardt, W.-D., Erdmann V.A., and Hartmann R.K. 1996. Rp-deoxy-phosphorothioate mod- ification interference experiments identify 2-OH groups in RNase P RNA that are cru- cial to tRNA binding. RNA 2: 1189–1198. Hardt W.-D., Schlegl J., Erdmann V.A., and Hartmann R.K. 1993. Role of the D arm and the anticodon arm in tRNA recognition by eubacterial and eukaryotic RNase P enzymes. Biochemistry 32: 13046–13053. Hardt W.-D., Warnecke J.M., Erdmann V.A., and Hartmann R.K. 1995. Rp-phosphoro- thioate modifications in RNase P RNA that interfere with tRNA binding. EMBO J. 14: 2935–2944. Harris M.E. and Pace N.R. 1995. Identification of phosphates involved in catalysis by the ribozyme RNase P RNA. RNA 1: 210–218. Harris M.E., Kazantsev A.V., Chen J.-L., and Pace N.R. 1997. Analysis of the tertiary structure of the ribonuclease P ribozyme-substrate complex by site-specific pho- toaffinity crosslinking. RNA 3: 561–576. Harris M.E., Nolan J.M., Malhotra A., Brown J.W., Harvey S.C., and Pace N.R. 1994. Use of photoaffinity crosslinking and molecular modeling to analyze the global structure of ribonuclease P RNA. EMBO J. 13: 3953–3963. Hartmann R.K., Heinrich J., Schlegl J., and Schuster H. 1995. Precursor of C4 antisense RNA of bacteriophages P1 and P7 is a substrate for RNase P of Escherichia coli. Proc. Natl. Acad. Sci. 92: 5822–5826. Herrmann B., Winqvist O., Mattsson J.G., and Kirsebom L.A. 1996. Differentiation of Chlamydia spp. by sequence determination and restriction cleavage of RNase P RNA genes. J. Clin. Microbiol. 34: 1897–1902. Himmelreich R., Hilbert H., Plagens H., Pirkl E., Li B.C., and Herrmann R. 1996. Complete sequence analysis of the genome of the bacterium Mycoplasma pneumo- niae. Nucleic Acids Res. 24: 4420–4449. Hollingsworth M.J. and Martin N.C. 1986. RNase P activity in the mitochondria of Saccharomyces cerevisiae depend on both mitochrondrion and nucleus-encoded com- ponents. Mol. Cell. Biol. 6: 1058–1064. Jarrous N., Eder P.S., Guerrier-Takada C., Hoog C., and Altman S. 1998. Autoantigenic properties of some protein subunits of catalytically active complexes of human ribonu- clease P. RNA 4: 407–417. Kahle D., Küst B., and Krupp G. 1993. Phosphorothioates in pre-tRNAs can change the specificities of RNases P or reduce the cleavage efficiencies. Biochimie 75: 955–962. Kahle D., Wehmeyer U., and Krupp G. 1990. Substrate recognition by RNase P and by the catalytic M1 RNA: Identification of possible contact points in pre-tRNAs. EMBO J. 9: 1929–1937. Kazakov S. and Altman S. 1991. Site-specific cleavage by metal ion cofactors and Ribonuclease P 377

inhibitors of M1 RNA, the catalytic subunit of RNase P from Escherichia coli. Proc. Natl. Acad. Sci. 88: 9193–9197. Keiler K.C., Waller R.H., and Sauer R.T. 1996. Role of a peptide tagging system in degradation of proteins synthesized from damaged messenger RNA. Science 271: 990–993. Kikuchi Y. and Suzuki-Fujita K. 1995. Synthesis and self-cleavage reaction of a chimeric molecule between RNase P-RNA and its model substrate. J. Biochem. 117: 197–200. Kikuchi Y., Sasaki-Tozawa N., and Suzuki K. 1993. Artificial self-cleaving molecules consisting of tRNA precursor and the catalytic RNA of RNase P. Nucleic Acids Res. 21: 4685–4689. Kirsebom L.A. 1995. RNase P—A ‘Scarlet Pimpernel.’ Mol. Microbiol. 17: 411–420. Kirsebom L.A. and Altman S. 1989. Reaction in vitro of some mutants of RNase P with wild-type and temperature-sensitive substrates. J. Mol. Biol. 207: 837–840. Kirsebom L.A. and Svärd S.G. 1992. The kinetics and specificity of cleavage by RNase P is mainly dependent on the structure of the amino acid acceptor stem. Nucleic Acids Res. 20: 425–432. ———. 1993. Identification of a region within M1 RNA of Escherichia coli RNase P important for the location of the cleavage site on a wild-type tRNA precursor. J. Mol. Biol. 231: 594–604. ———. 1994. Base pairing between Escherichia coli RNase P RNA and its substrate. EMBO J. 13: 4870–4876. Kleineidam R.G., Pitulle C., Sproat B., and Krupp G. 1993. Efficient cleavage of pre- tRNAs by E. coli RNase P RNA requires the 2-hydroxyl of the ribose at the cleavage site. Nucleic Acids Res. 21: 1097–1101. Knap A.K., Wesolowski D., and Altman S. 1990. Protection from chemical modification of nucleotides in complexes of M1 RNA, the catalytic subunit of RNase P from E. coli, and tRNA precursors. Biochimie 72: 779–790. Komine Y., Kitabatake M., Yokogawa T., and Nishikawa K. 1994. A tRNA-like structure is present in 10Sa RNA, a small stable RNA from E. coli. Proc. Natl. Acad. Sci. 91: 9223–9277. Kufel J. and Kirsebom L.A. 1994. Cleavage site selection by M1 RNA, the catalytic sub- unit of Escherichia coli RNase P, is influenced by pH. J. Mol. Biol. 244: 511–521. ———. 1996a. Different cleavage sites are aligned differently in the active site of M1 RNA, the catalytic subunit of Escherichia coli RNase P. Proc. Natl. Acad. Sci. 93: 6085–6090. ———. 1996b. Residues in Escherichia coli RNase P RNA important for cleavage site selection and divalent metal ion binding. J. Mol. Biol. 263: 685–698. Kunst F., Ogasawara N., Moszer I., Albertini A.M., Alloni G., Azevedo V., Bertero M.G., Bessieres P., Bolotin A., Borchert S., Borriss R., Boursier L., Brans A., Braun M., Brignell S.C., Bron S., Brouillet S., Bruschi C.V., Caldwell B., Capuano V., Carter N.M., Choi S.K., Codani J.J., Connerton I.F., Danchin A., et al. 1997. The complete genome sequence of the gram-positive bacterium Bacillus subtilis. Nature 390: 249–256. Lang B.F., Burger G., O’Kelly C.J., Cedergren R., Golding G.B., Lemieux C., Sankoff D., Turmel M., and Grey M.W. 1997. An ancestral mitochondrial DNA resembling a eubacterial genome in miniature. Nature 387: 493–496. Lee Y., Kindelberger D.W., Lee J-Y., McClennen S., Chamberlain J., and Engelke D.R. 1997. Nuclear pre-tRNA terminal structure and RNase P recognition. RNA 3: 175–185. LeGrandeur T.E., Huttenhofer A., Noller N.F., and Pace N.R. 1994. Phylogenetic com- 378 S. Altman and L.A. Kirsebom

parative chemical footprint analysis of the interaction between ribonuclease P RNA and tRNA. EMBO J. 13: 3945–3952. Levinger L., Bourne R., Kolla S., Cylin E., Russell K., Wang X., and Mohan A. 1997. Processing kinetics and minimal substrates for Drosophila RNase P and 3-tRNase. Nucleic Acids Symp. Ser. 36: 78–82. Li K. and Williams R.S. 1995. Cloning and characterization of three new murine genes encoding short homologues of RNase P RNA. J. Biol. Chem. 42: 25281–25285. Li Y. and Altman S. 1996. Cleavage by RNase P of gene N mRNA reduces bacteriophage 1 burst size. Nucleic Acids Res. 24: 835–842. Li Y., Guerrier-Takada C., and Altman S. 1992. Targeted cleavage of mRNA in vitro by RNase P from Escherichia coli. Proc. Natl. Acad. Sci. 89: 3185–3189. Liu F. and Altman S. 1994. Differential evolution of substrates for an RNA enzyme in the presence and absence of its protein cofactor. Cell 77: 1093–1100. ———. 1995. Inhibition of viral gene expression by the catalytic RNA subunit of RNase P from Escherichia coli. Genes Dev. 9: 471–480. ———. 1996. Requirements for cleavage by a modified RNase P of a small model sub- strate. Nucleic Acids Res. 24: 2690–2696. Loria A. and Pan T. 1997. Recognition of the T stem-loop of a pre-tRNA substrate by the ribozyme from Bacillus subtilis ribonuclease P. Biochemistry 36: 6317–6325. Lygerou Z., Pluk H., van Venrooij W.J., and Seraphin B. 1996. hPop1: An autoantigenic protein subunit shared by the human RNase P and RNase MRP ribonucleoproteins. EMBO J. 15: 5936–5948. Maizels N. and Weiner A.M. 1993. The genomic tag hypothesis: Modern viruses as molecular fossils of ancient strategies for genomic replication. In The RNA world (ed. R.F. Gesteland and J.F. Atkins), pp. 577–602. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, New York. Martin N.C. and Lang B.F. 1997. Mitochondrial RNase P: The RNA family grows. Nucleic Acids Symp. Ser. 36: 42–44. Massire C., Jaeger L., and Westhof E. 1998. Derivation of the three-dimensional architec- ture of bacterial ribonuclease P RNAs from comparative sequence analysis. J. Mol. Biol. 279: 773–793. McClain W.H., Guerrier-Takada C., and Altman S. 1987. Model substrates for an RNA enzyme. Science 238: 527–530. Met Meinnel T. and Blanquet S. 1995. Maturation of pre-tRNA f by Escherichia coli RNase P is specified by a guanosine of the 5-flanking sequence. J. Biol. Chem. 270: 15908–15914. Muto A., Sato M., Tadaki T., Fukushima M., Ushida C., and Himeno H. 1996. Structure and function of 10Sa RNA: trans-translation system. Biochimie 78: 985–991. Nolan J.M., Burke D.H., and Pace N.R. 1993. Circulary permutated tRNAs as specific photoaffinity probes of ribonuclease P RNA structure. Science 261: 762–765. Oh B-K. and Pace N.R. 1994. Interaction of the 3-end of tRNA with ribonuclease P RNA. Nucleic Acids Res. 22: 4087–4094. Pan T. 1995. Higher order folding and domain analysis of the ribozyme B. subtilis ribonu- clease P. Biochemistry 34: 902–909. Pan T. and Jakacka M. 1996. Multiple substrate binding sites in the ribozyme from Bacillus subtilis RNase P. EMBO J. 15: 2249–2255. Peck-Miller K.A. and Altman S. 1991. Kinetics of the processing of the precursor to 4.5 S Ribonuclease P 379

RNA, a naturally occurring substrate for RNase P from Escherichia coli. J. Mol. Biol. 221: 1–5. Perreault J-P. and Altman S. 1992. Important 2-hydroxyl groups in model substrates for M1 RNA, the catalytic subunit of RNase P from Escherichia coli. J. Mol. Biol. 226: 399–409. ———. 1993. Pathway of activation by magnesium ions of substrates for the catalytic subunit of RNase P from Escherichia coli. J. Mol. Biol. 230: 750–756. Poritz M.A., Bernstein H.D., Strub K., Zopf D., Wilhelm H., and Walter P. 1990. An E. coli ribonucleoprotein containing 4.5S RNA resembles mammalian signal recognition particle. Science 250: 1111–1117. Reich C., Olsen G.J., Pace B., and Pace N.R. 1988. Role of the protein moiety of RNase P, a ribonucleoprotein enzyme. Science 239: 178–181. Seidman J.G. and McClain W.H. 1975. Three steps in conversion of large precursor RNA into serine and proline transfer RNAs. Proc. Natl. Acad. Sci. 72: 1491–1495. Siegal R.W., Banta A.B., Haas E.S., Brown J.W., and Pace N.R. 1996. Mycoplasma fer- mentans simplifies our view of the catalytic core of ribonuclease P RNA. RNA 2: 452–462. Smith D. and Pace N.R. 1993. Multiple magnesium ions in the ribonuclease P reaction mechanism. Biochemistry 32: 5273–5281. Stolc V. and Altman S. 1997. Rpp1, an essential protein subunit of nuclear RNase P required for processing of precursor tRNA and 35S precursor rRNA in Saccharomyces cerevisiae. Genes Dev. 11: 2926–2937. Stolc V., Katz A., and Altman S. 1998. Rpp2, an essential protein subunit of nuclear RNase P is required for processing of precursor tRNAs and 35S precursor rRNA in Saccharomyces cerevisiae. Proc. Natl. Acad. Sci. 95: 6716–6721. Svärd S.G. and Kirsebom L.A. 1992. Several regions of a tRNA precursor determine the Escherichia coli RNase P cleavage site. J. Mol. Biol. 227: 1019–1031. ———. 1993. Determinants of Escherichia coli RNase P cleavage site selection: A detailed in vitro and in vivo analysis. Nucleic Acids Res. 21: 427–434. Svärd S.G., Kagardt U., and Kirsebom L.A. 1996. Phylogenetic comparative mutational analysis of the base-pairing between RNase P RNA and its substrate. RNA 2: 463–472. Tallsjö A. and Kirsebom L.A. 1993. Product release is a rate-limiting step during cleavage by the catalytic RNA subunit of Escherichia coli RNase P. Nucleic Acids Res. 21: 51–57. Tallsjö A., Kufel J., and Kirsebom L.A. 1996. Interaction between Escherichia coli RNase P RNA and the discriminator base results in slow product release. RNA 2: 299–307. Thurlow D.L., Shilowski D., and Marsh T.L. 1991. Nucleotides in precursor tRNAs that are required intact for catalysis by RNase P RNAs. Nucleic Acids Res. 19: 885–891. Tollervey D. 1995. Genetic and biochemical analyses of yeast RNase MRP. Mol. Biol. Rep. 22: 75–79. Tomb J.F., White O., Kerlavage A.R., Clayton R.A., Sutton G.G., Fleischmann R.D., Ketchum K.A., Klenk H.P., Gill S., Dougherty B.A., Nelson K., Quackenbush J., Zhou L., Kirkness E.F., Peterson S., Loftus B., Richardson D., Dodson R., Khalak H.G., Glodek A., McKenney K., Fitzegerald L.M., Lee N., Adams M.D., Venter J.C., et al. 1997. The complete genome sequence of the gastric pathogen Helicobacter pylori. Nature 388: 539–547. 380 S. Altman and L.A. Kirsebom

Vioque A. 1997. The RNase P RNA from cyanobacteria: Short tandemly repeated repeti- tive (STRR) sequences are present within the RNase P RNA gene in heterocyst- forming cyanobacteria. Nucleic Acids Res. 25: 3471–3477. Wang M.J., Davis N.W., and Gegenheimer P. 1988. Novel mechanisms for maturation of transfer RNA precursors. EMBO J. 7: 1567–1574. Warnecke J.M., Fürste J.P., Hardt W.-D., Erdmann V.A., and Hartmann R.K. 1996. Ribonuclease P (RNase P) RNA is converted to a Cd2+-ribozyme by a single Rp- phosphorothioate modification in the precursor tRNA at the RNase P cleavage site. Proc. Natl. Acad. Sci. 93: 8924–8928. Werner M., Rosa E., and George S. 1997. Design of short external guide sequences (EGSs) for cleavage of target molecules with RNase P. Nucleic Acids. Symp. Ser. 36: 19–21. Westhof E. and Altman S. 1994. Three dimensional working model of M1 RNA, the cat- alytic RNA subunit of ribonuclease P from Escherichia coli. Proc. Nat. Acad. Sci. 91: 5133–5137. Westhof E., Wesolowski D., and Altman S. 1996. Mapping in three dimensions regions in a catalytic RNA protected from attack by an Fe(II)-EDTA reagent. J. Mol. Biol. 258: 600–613. Wise C. and Martin N.C. 1991. Dramatic size variation of yeast mitochondrial RNAs sug- gests that RNase P RNAs can be quite small. J. Biol. Chem. 266: 19154–19157. Yuan Y. and Altman S. 1995. Substrate recognition by human RNase P identification of small, model substrates for the enzyme. EMBO J. 14: 159–168. Yuan Y., Hwang E.-S., and Altman S. 1992. Targeted cleavage of mRNA by human RNase P. Proc. Natl. Acad. Sci. 89: 8006–8010. Yuan Y., Tan E., and Reddy R. 1991. The 40-kilodalton to autoantigen associates with nucleotides 21 to 64 of human mitochondrial RNA processing/7-2 RNA in vitro. Mol. Cell. Biol. 11: 5266–5274. Zito K., Huttenhofer A., and Pace N.R. 1993. Lead-catalyzed cleavage of ribonuclease P RNA as a probe for integrity of tertiary structure. Nucleic Acids Res. 21: 5916–5920.