11.02.2013 Views

Programmed Cell Death by hok/sok of Plasmid - Department of Plant ...

Programmed Cell Death by hok/sok of Plasmid - Department of Plant ...

Programmed Cell Death by hok/sok of Plasmid - Department of Plant ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Programmed</strong> <strong>Cell</strong> <strong>Death</strong> <strong>by</strong> <strong>hok</strong>/<strong>sok</strong> <strong>of</strong> <strong>Plasmid</strong> R1:<br />

Processing at the <strong>hok</strong> mRNA 3 0 -end Triggers<br />

Structural Rearrangements that Allow Translation and<br />

Antisense RNA Binding<br />

Thomas Franch 1 , Alexander P. Gultyaev 2 and Kenn Gerdes 1 *<br />

1 <strong>Department</strong> <strong>of</strong> Molecular<br />

Biology, Odense University<br />

Campusvej 55, DK-5230<br />

Odense M, Denmark<br />

2 Leiden Institute <strong>of</strong> Chemistry<br />

Leiden University, P.O. Box<br />

9502, 2300 RA, Leiden<br />

The Netherlands<br />

*Corresponding author<br />

Introduction<br />

The <strong>hok</strong>/<strong>sok</strong> system <strong>of</strong> plasmid R1 codes for two<br />

RNAs, <strong>hok</strong> mRNA and Sok antisense RNA (Gerdes<br />

et al., 1988, 1990). The mRNA encodes two reading<br />

frames, denoted <strong>hok</strong> (host killing) and mok (modulation<br />

<strong>of</strong> killing), that overlap extensively (Figure 1).<br />

Translation <strong>of</strong> <strong>hok</strong> is coupled to translation <strong>of</strong> mok<br />

(Thisted & Gerdes, 1992). Sok-RNA represses<br />

translation <strong>of</strong> mok via binding to the translational<br />

initiation region (TIR) <strong>of</strong> mok (Thisted et al., 1994a).<br />

Therefore, Sok-RNA regulates <strong>hok</strong> translation<br />

Abbreviations used: SD, Shine-Dalgarno; ucb,<br />

upstream complementary box; dcb, downstream<br />

complementary box; tac, translational activator; TIR,<br />

translational initiation region; DMS, dimethyl sulphate;<br />

wt, wild-type; IPTG, isopropyl-b,Dthiogalactopyranoside.<br />

J. Mol. Biol. (1997) 273, 38±51<br />

The <strong>hok</strong>/<strong>sok</strong> locus <strong>of</strong> plasmid R1 mediates plasmid stabilization <strong>by</strong> killing<br />

<strong>of</strong> plasmid-free cells. The locus speci®es two RNAs, <strong>hok</strong> mRNA and Sok<br />

antisense RNA. The post-segregational killing mediated <strong>by</strong> <strong>hok</strong>/<strong>sok</strong> is governed<br />

<strong>by</strong> a complicated control mechanism that involves both post-transcriptional<br />

inhibition <strong>of</strong> translation <strong>by</strong> Sok-RNA and activation <strong>of</strong> <strong>hok</strong><br />

translation <strong>by</strong> mRNA 3 0 processing. Sok-RNA inhibits translation <strong>of</strong> a<br />

reading frame (mok) that overlaps with <strong>hok</strong>, and translation <strong>of</strong> <strong>hok</strong> is<br />

coupled to translation <strong>of</strong> mok. In the inactive full-length <strong>hok</strong> mRNA, the<br />

translational activator element at the mRNA 5 0 -end (tac) is sequestered <strong>by</strong><br />

the fold-back-inhibitory element located at the mRNA 3 0 -end (fbi). The 5 0<br />

to 3 0 pairing locks the RNA in an inert con®guration in which the SD mok<br />

and Sok-RNA target regions are sequestered. Here we show that the 3 0<br />

processing leads to major structural rearrangements in the mRNA 5 0 -end.<br />

The structure <strong>of</strong> the refolded RNA explains activation <strong>of</strong> translation and<br />

antisense RNA binding. The refolded RNA contains an antisense RNA<br />

target stem-loop that presents the target nucleotides in a single-stranded<br />

conformation. The stem <strong>of</strong> the target hairpin contains SD mok and AUG mok<br />

in a paired con®guration. Using toeprinting analysis, we show that this<br />

pairing keeps SD mok in an accessible con®guration. Furthermore, a mutational<br />

analysis shows that an internal loop in the target stem is prerequisite<br />

for ef®cient translation and antisense RNA binding.<br />

# 1997 Academic Press Limited<br />

Keywords: Sok; <strong>hok</strong>; antisense RNA; translational initiation; RNA<br />

rearrangements<br />

indirectly through mok. The 64 nt Sok-RNA consists<br />

<strong>of</strong> a stem±loop with an 11 nucleotide 5 0<br />

single-stranded extension (Figure 2A). The 5 0 extension<br />

is complementary to the part <strong>of</strong> the mok TIR<br />

that we denote <strong>sok</strong>T 0 (Figure 2) and is responsible<br />

for the initial recognition reaction between the<br />

RNAs (Thisted et al., 1994a). After initial recognition,<br />

more extensive duplex formation progresses<br />

from the Sok-RNA 5 0 -end <strong>by</strong> a zippering-like<br />

mechanism. Subsequently, such RNA duplexes are<br />

cleaved <strong>by</strong> RNase III (Gerdes et al., 1992). Thus,<br />

binding <strong>of</strong> Sok-RNA to <strong>hok</strong> mRNA occurs <strong>by</strong> a<br />

one-step binding mechanism similar to the RNA-<br />

IN/RNA-OUT pairing in Tn10 (Kittle et al., 1989;<br />

Case et al., 1989).<br />

The complex RNA metabolism prerequisite for<br />

activation <strong>of</strong> <strong>hok</strong> translation in plasmid-free cells is<br />

described in detail in the accompanying paper<br />

(Gultyaev et al., 1997) and is brie¯y outlined<br />

0022±2836/97/410038±14 $25.00/0/mb971294 # 1997 Academic Press Limited


Structure and Function <strong>of</strong> <strong>hok</strong> mRNA <strong>of</strong> <strong>Plasmid</strong> R1 39<br />

Figure 1. Overview <strong>of</strong> the structural elements <strong>of</strong> the <strong>hok</strong>/<br />

<strong>sok</strong> system from plasmid R1. FL-1 and FL-2 denote the<br />

full-length <strong>hok</strong> mRNA-1 and mRNA-2, respectively. TR<br />

denotes the 3 0 truncated <strong>hok</strong> mRNA. Other abbreviations<br />

used: <strong>hok</strong>, host killing; mok, modulation <strong>of</strong> killing; Sok,<br />

suppression <strong>of</strong> killing;, <strong>sok</strong>T, Sok-RNA target; fbi, foldback-inhibition;<br />

and tac, translational activation.<br />

below. We showed previously that <strong>hok</strong> mRNA<br />

exists in two variants in vivo (Gerdes et al., 1990;<br />

Thisted et al., 1994a). A truncated, translationally<br />

active mRNA is generated <strong>by</strong> slow exonucleolytic<br />

removal <strong>of</strong> 40 nt from the 3 0 -end <strong>of</strong> the full-length<br />

mRNA. Thus, the full-length mRNA acts as a reservoir<br />

from which the active mRNA is continuously<br />

generated. In plasmid-carrying cells, the presence<br />

<strong>of</strong> Sok-RNA prevents translation <strong>of</strong> the truncated<br />

mRNA. However, in plasmid-free cells, in which<br />

Sok-RNA has decayed, the continuous slow processing<br />

<strong>of</strong> the full-length mRNA leads to accumulation<br />

<strong>of</strong> the translatable truncated mRNA.<br />

Eventually, this leads to synthesis <strong>of</strong> the toxic Hok<br />

protein, and killing <strong>of</strong> the plasmid-free cells<br />

ensues.<br />

Recently, we showed that the stable, inactive <strong>hok</strong><br />

mRNA is compactedly folded with only a few<br />

single-stranded regions, apart from loops (Franch<br />

& Gerdes, 1996). A genetic analysis established<br />

that the full-length mRNA contains a peculiar 5 0 to<br />

3 0 long-range pairing in which the very ®rst<br />

nucleotides pair with the very last ones (Figure 2B).<br />

This interaction is prerequisite for the post-segregational<br />

killing mechanism, since its disruption (<strong>by</strong><br />

mutation) leads to mRNA instability and depletion<br />

<strong>of</strong> the activatable pool <strong>of</strong> full-length mRNA<br />

(Franch & Gerdes, 1996). The 3 0 -element responsible<br />

for inhibition <strong>of</strong> translation and antisense<br />

RNA binding was called fbi (fold-back-inhibition).<br />

The 5 0 element that pairs with fbi was denoted tac<br />

(translational activation). The tac element is<br />

required for activation <strong>of</strong> translation (Franch &<br />

Gerdes, 1996).<br />

In full-length <strong>hok</strong> mRNA, the SD mok element is<br />

sequestered <strong>by</strong> an upstream sequence that we<br />

denote ucb (upstream complementary box). The<br />

ucb element is shown in red in Figure 2. Previously<br />

we suggested that pairing between SD mok and ucb<br />

prevents ribosome entry at SD mok and there<strong>by</strong> inhibits<br />

translation <strong>of</strong> mok (and <strong>hok</strong>; Nielsen & Gerdes,<br />

1995). A nearly perfect repeat <strong>of</strong> the 9 nt ucb<br />

element is located 58 nt downstream <strong>of</strong> ucb<br />

(Figure 2). This element, which pairs with the<br />

translation initiation region <strong>of</strong> <strong>hok</strong> (TIR <strong>hok</strong>) in fulllength<br />

<strong>hok</strong> mRNA, was denoted dcb (downstream<br />

complementary box). This nomenclature is introduced<br />

to illustrate the complex structural<br />

rearrangements that occur in <strong>hok</strong> mRNA (see<br />

below). As a consequence <strong>of</strong> their repetitive nature,<br />

the ucb and dcb elements are both complementary<br />

to SD mok (compare Figure 2B and D).<br />

The structure <strong>of</strong> the antisense RNA target in <strong>hok</strong><br />

mRNA is yet unknown. However, the genetic<br />

algorithm calculations and phylogenetic comparisons<br />

presented in the accompanying paper suggest<br />

that the exonucleolytic removal <strong>of</strong> fbi triggers<br />

major structural rearrangements at the mRNA 5 0 -<br />

end. More precisely, the computer-simulated RNA<br />

foldings predict that in truncated mRNA tac pairs<br />

with ucb, and SD mok with dcb. The latter interaction<br />

is involved in the formation <strong>of</strong> an antisense RNA<br />

target hairpin (Figure 2D). The target hairpin is<br />

topped <strong>by</strong> a 7 nt loop complementary to the very<br />

5 0 -end <strong>of</strong> Sok-RNA.<br />

Here, we investigate the biological properties<br />

and secondary structures <strong>of</strong> the full-length and<br />

truncated <strong>hok</strong> mRNAs. We show that the binding<br />

<strong>of</strong> Sok-RNA to full-length <strong>hok</strong> mRNA is negligible<br />

both in vivo and in vitro. Using in vitro toeprint<br />

analyses, we show that the mok TIR in full-length<br />

<strong>hok</strong> mRNA is unable to bind ribosomal 30 S subunits,<br />

in accordance with the observation that<br />

full-length <strong>hok</strong> mRNA is not translated in vivo.<br />

Most importantly, we obtain ®rm evidence for<br />

the postulated structural rearrangements triggered<br />

<strong>by</strong> 3 0 processing <strong>of</strong> the full-length mRNA. In the<br />

truncated mRNA, tac was found to pair with ucb,<br />

and SD mok with dcb, thus con®rming the presence<br />

<strong>of</strong> the tac and target hairpins. A mutational analysis<br />

shows that the structure <strong>of</strong> the target hairpin<br />

is crucial for rapid antisense RNA binding and<br />

for translation <strong>of</strong> mok, and that even modest<br />

changes in its structure have pr<strong>of</strong>ound effects on<br />

these parameters. We further show that the pairing<br />

between SD mok and dcb is required to maintain<br />

truncated <strong>hok</strong> mRNA in a translatable<br />

con®guration.<br />

Results<br />

Sok antisense RNA binds to truncated but not<br />

to full-length <strong>hok</strong> mRNA in vivo<br />

We have previously examined the metabolism <strong>of</strong><br />

the <strong>hok</strong> mRNAs in vivo (Gerdes et al., 1990, 1992;<br />

Thisted & Gerdes, 1992; Thisted et al., 1994a). Indirect<br />

evidence from Northern analyses suggested<br />

that Sok-RNA binds to truncated <strong>hok</strong> mRNA much<br />

more ef®ciently than to the full-length RNA. To<br />

test this inference more directly, we constructed a<br />

plasmid vector carrying a <strong>hok</strong>/<strong>sok</strong> system in which<br />

P <strong>hok</strong> was precisely replaced <strong>by</strong> the strong, LacI<br />

repressed P A1/04 promotor (Lanzer & Bujard, 1988).<br />

In this system, synthesis <strong>of</strong> <strong>hok</strong> mRNA is halted <strong>by</strong><br />

the removal <strong>of</strong> isopropyl-b,D-thiogalactopyranoside<br />

(IPTG) from the growth-medium. Thus, synthesis


40 Structure and Function <strong>of</strong> <strong>hok</strong> mRNA <strong>of</strong> <strong>Plasmid</strong> R1<br />

<strong>of</strong> <strong>hok</strong> mRNA could be blocked without the<br />

addition <strong>of</strong> rifampicin.<br />

The in vivo processing pattern <strong>of</strong> the <strong>hok</strong> mRNAs<br />

was analysed <strong>by</strong> Northern blotting in the presence<br />

(<strong>sok</strong> ‡ ) or absence (<strong>sok</strong> ) <strong>of</strong> antisense RNA. In its<br />

presence, two <strong>hok</strong> mRNAs, denoted full-length<br />

mRNA-1 and mRNA-2, were observed (Figure 3A,<br />

®rst panel). Both mRNAs were quite stable with a<br />

half-life <strong>of</strong> approximately 20 minutes. Previous<br />

analyses using rifampicin to inhibit transcription<br />

yielded half-lives <strong>of</strong> the order <strong>of</strong> 20 to 40 minutes<br />

(Gerdes et al., 1990, 1992; Thisted et al., 1994a). <strong>hok</strong><br />

Figure 2. Secondary structures <strong>of</strong> Sok-RNA and <strong>hok</strong> mRNAs. A, Secondary structure <strong>of</strong> Sok-RNA. B, Secondary structure<br />

<strong>of</strong> full-length <strong>hok</strong> mRNA. C, Con®guration <strong>of</strong> non-refolded truncated <strong>hok</strong> mRNA as predicted <strong>by</strong> computerassisted<br />

structure calculations. D, Putative truncated, refolded mRNA supported <strong>by</strong> the structure-probing analyses<br />

presented in Figure 4 and the RNase H cleavage pattern shown in Figure 5. Shine-Dalgarno and start codons <strong>of</strong> the<br />

mok and <strong>hok</strong> reading frames are shown in blue, and the ucb and dcb repetitive sequences are shown in red. C74 in<br />

ucb, marked in green, emphasizes a heterogeneity between the ucb and dcb elements. Thin numbered lines mark the<br />

nucleotides complementary to the three DNA oligonucleotides used in the RNase H cleavage assay. The antisense<br />

RNA target that is complementary to the 5 0 -tail <strong>of</strong> Sok-RNA is denoted <strong>sok</strong>T 0 (B). The sequence and structure <strong>of</strong><br />

super-tac truncated RNA is shown below the wt tac/ucb pairing in D. The non-refolded (C) and refolded (D) con®gurations<br />

are proposed to exist in a dynamic equilibrium.


Structure and Function <strong>of</strong> <strong>hok</strong> mRNA <strong>of</strong> <strong>Plasmid</strong> R1 41<br />

Figure 3. A, Effect <strong>of</strong> Sok antisense RNA on the stability<br />

and processing pattern <strong>of</strong> <strong>hok</strong> mRNAs as shown <strong>by</strong><br />

Northern blotting. The ®rst panel shows <strong>hok</strong> mRNAs<br />

after arrest <strong>of</strong> <strong>hok</strong> transcription in the presence <strong>of</strong> Sok-<br />

RNA, the second panel shows <strong>hok</strong> mRNAs after arrest<br />

<strong>of</strong> transcription in the absence <strong>of</strong> Sok-RNA. <strong>Cell</strong>s<br />

(CSH50 containing either pKG1540 (<strong>sok</strong> ‡ ) or pKG1541<br />

(<strong>sok</strong> )) were grown in LB medium containing 1 mM<br />

IPTG to an A 450 <strong>of</strong> ca 0.2, harvested, and resuspended in<br />

prewarmed medium without IPTG. <strong>Cell</strong> samples for<br />

total RNA preparation were taken at the intervals indicated.<br />

The mRNAs contained stop codon mutations in<br />

the <strong>hok</strong> gene to prevent detrimental <strong>hok</strong> expression (from<br />

Thisted & Gerdes, 1992). B, In vitro binding assay<br />

between 32 P-labelled Sok-RNA and unlabelled ( 3 H) truncated<br />

<strong>hok</strong> mRNA (®rst panel) and full-length <strong>hok</strong><br />

mRNA-1 (second panel). Labelled antisense RNA was<br />

added to a tenfold excess <strong>of</strong> target RNA. For details, see<br />

Materials and Methods.<br />

mRNA-2 is generated from <strong>hok</strong> mRNA-1 <strong>by</strong> an as<br />

yet unknown mechanism (Nikolaj Dam Mikkelsen,<br />

unpublished results). The absence <strong>of</strong> Sok-RNA<br />

caused the appearance <strong>of</strong> a third band corresponding<br />

to truncated <strong>hok</strong> mRNA (Figure 3A, second<br />

panel). This RNA is generated <strong>by</strong> the exonucleolytic<br />

removal <strong>of</strong> 40 nt at the 3 0 -end <strong>of</strong> <strong>hok</strong> mRNA-2<br />

(Gerdes et al., 1990). Comparison <strong>of</strong> the two panels<br />

<strong>of</strong> Figure 3A yields valuable new information on<br />

the in vivo metabolism <strong>of</strong> <strong>hok</strong> mRNA: ®rst, the<br />

absence <strong>of</strong> Sok-RNA has no in¯uence on the<br />

amounts or the stabilities <strong>of</strong> the two full-length<br />

species. Since antisense RNA binding leads to<br />

duplex formation and rapid RNase III cleavage,<br />

this observation indicates that Sok-RNA does not<br />

bind to <strong>hok</strong> mRNA-1 or mRNA-2 at a signi®cant<br />

rate. Secondly, the absence <strong>of</strong> Sok-RNA resulted in<br />

the accumulation <strong>of</strong> the truncated mRNA. Since<br />

this RNA is absent from the ®rst panel, this observation<br />

indicates that Sok-RNA binds avidly to<br />

truncated <strong>hok</strong> mRNA and there<strong>by</strong> confers its rapid<br />

decay via RNase III-mediated hydrolysis. Taken<br />

together, these results show that Sok-RNA binds<br />

full-length <strong>hok</strong> mRNA at a negligible rate in vivo,<br />

whereas the truncated version <strong>of</strong> <strong>hok</strong> mRNA is scavenged<br />

rapidly.<br />

Sok-RNA binds 100-fold more rapidly to<br />

truncated than to full-length <strong>hok</strong> mRNA<br />

in vitro<br />

The above results indicate that Sok-RNA binds<br />

much more rapidly to truncated than to full-length<br />

<strong>hok</strong> mRNA in vivo. Using a standard in vitro binding<br />

assay, we showed previously that full-length<br />

<strong>hok</strong> mRNA bound Sok-RNA with an approximately<br />

30% reduced rate (Thisted et al., 1994b). To<br />

resolve this apparent discrepancy, we repeated the<br />

in vitro binding assay using a more physiologically<br />

correct binding buffer (i.e. 200 mM potassium glutamate<br />

was included). The resulting gel-shift experiment<br />

is shown in Figure 3B. As seen, Sok-RNA<br />

bound readily to the truncated RNA (Figure 3A).<br />

However, binding to the full-length RNA was<br />

inhibited (Figure 3B). The apparent second-order<br />

binding-rate constants (K app) were estimated to<br />

be 3 10 5 M 1 s 1 for truncated and less than<br />

3 10 3 M 1 s 1 for full-length <strong>hok</strong> mRNA, corresponding<br />

to a more than 100-fold difference in<br />

binding-rates (for calculations, see Materials and<br />

Methods). This result suggests that the <strong>sok</strong>T 0 region<br />

exists in different con®gurations in the two RNAs<br />

(see Figure 2B and D). Now we can address this<br />

important question experimentally.<br />

Secondary structure analyses <strong>of</strong> fulllength,<br />

truncated and super-tac truncated<br />

<strong>hok</strong> mRNAs<br />

Mutational analyses, computer calculations and<br />

phylogenetic comparisons suggested that translational<br />

activation <strong>of</strong> <strong>hok</strong> mRNA involves a substantial<br />

structural rearrangement at the mRNA 5 0 -end,<br />

and that this rearrangement is triggered <strong>by</strong> the 3 0<br />

processing <strong>of</strong> full-length <strong>hok</strong> mRNA (Franch &<br />

Gerdes, 1996; Gultyaev et al., 1997, accompanying<br />

paper). More speci®cally, our analyses predicted<br />

that in truncated RNA, tac preferentially pairs with<br />

ucb there<strong>by</strong> leading to formation <strong>of</strong> the extended<br />

tac-stem (see Figure 2D). Although a large internal<br />

loop destabilizes the tac/ucb interaction, it contains<br />

the possibility <strong>of</strong> non-canonical GA/AG base-pairing<br />

(Turner & Bevilacqua, 1993), which may add to<br />

the energy <strong>of</strong> the tac stem. Most importantly,<br />

the 5 0 refolding was accompanied <strong>by</strong> structural<br />

rearrangements further downstream in the molecule.<br />

Thus, a stem±loop that we coin the antisense


42 Structure and Function <strong>of</strong> <strong>hok</strong> mRNA <strong>of</strong> <strong>Plasmid</strong> R1<br />

Figure 4. Secondary structure probing <strong>by</strong> DMS (®rst panel), RNase T 2 (second panel) and RNase V 1 (third panel) <strong>of</strong><br />

full-length, truncated and super-tac truncated mRNA. The sequence and secondary structure <strong>of</strong> the super-tac mutation<br />

are shown in Figure 2D. Modi®cation/cleavage bands mentioned in the text are marked <strong>by</strong> arrowheads.<br />

RNA target hairpin was formed (Figure 2D and<br />

the accompanying paper, Gultyaev et al., 1997).<br />

Construction <strong>of</strong> a ``perfect'' tac stem <strong>by</strong> the introduction<br />

<strong>of</strong> the super-tac mutation (see Figure 2D)<br />

increased translation <strong>of</strong> the truncated <strong>hok</strong> mRNA<br />

tw<strong>of</strong>old (Franch & Gerdes, 1996). The super-tac<br />

mutation greatly increases the stability <strong>of</strong> the tac<br />

stem. Consequently, the presence <strong>of</strong> super-tac predictably<br />

would lead to refolding <strong>of</strong> the entire<br />

population <strong>of</strong> <strong>hok</strong> mRNA molecules present in a<br />

pool <strong>of</strong> in vitro prefolded RNA. This inference is in<br />

accordance with the increased translation rate <strong>of</strong><br />

super-tac truncated RNA.<br />

In order to verify the predicted secondary structure<br />

transitions, structure probing analyses were<br />

conducted on full-length, truncated and super-tac<br />

truncated RNAs (Figure 4). Chemical modi®cation<br />

<strong>by</strong> dimethyl sulphate (modi®es N1 <strong>of</strong> unpaired<br />

A > N3 <strong>of</strong> unpaired C), cleavage <strong>by</strong> RNase T 2<br />

(cleaves 3 0 <strong>of</strong> unpaired nucleotides) and RNase V 1<br />

(cleaves paired and stacked nucleotides) were<br />

detected <strong>by</strong> primer extension using the <strong>hok</strong>1 primer.<br />

For convenience, only the regulatory region<br />

covering nucleotides 70 to 160 is shown in<br />

Figure 4. However, the entire mRNA 5 0 -ends<br />

were also extensively analysed during this work<br />

(data not shown).<br />

Full-length <strong>hok</strong> mRNA<br />

In the full-length mRNA, the proposed position<br />

<strong>of</strong> the loop in the ucb/SD mok stem±loop structure


Structure and Function <strong>of</strong> <strong>hok</strong> mRNA <strong>of</strong> <strong>Plasmid</strong> R1 43<br />

was supported <strong>by</strong> T 2 cuts in the region U89-U91<br />

and DMS modi®cation at C92. The DMS modi®cation<br />

at A81 supports the internal loop at the top<br />

<strong>of</strong> the stem. Furthermore, V 1 cuts at 110 to 112<br />

could support the double-stranded bottom stem.<br />

Three A-C mismatches (C74-A107, A78-C103 and<br />

A82-C100) were predicted in the ucb/SD mok stem.<br />

No signi®cant modi®cation <strong>of</strong> these nucleotides<br />

was observed, probably due to the formation <strong>of</strong><br />

A-C reverse wobble pairings (Saenger, 1984). The<br />

modi®cations at C115 and A117, and to a lesser<br />

extent at C113 and C114, suggest that this region is<br />

at least partially single-stranded. The secondary<br />

structure probing <strong>of</strong> the full-length <strong>hok</strong> mRNA is<br />

consistent with the structure presented in Figure 2B<br />

in which <strong>sok</strong>T 0 is in a double-stranded conformation.<br />

Wild-type and super-tac truncated <strong>hok</strong> mRNAs<br />

The wild-type and super-tac truncated mRNAs<br />

exhibited nearly identical modi®cation/cleavage<br />

patterns and comparison with that <strong>of</strong> the fulllength<br />

RNA revealed important differences. Cleavage<br />

<strong>by</strong> T 2 at A78 and A81-A82 in the truncated<br />

mRNA suggests the presence <strong>of</strong> a single-stranded<br />

region not present in the full-length transcript.<br />

These cuts appeared enhanced in super-tac mRNA.<br />

In the SD mok region signi®cant DMS modi®cations<br />

were observed in super-tac mRNA compared to the<br />

weak probing <strong>of</strong> the wt truncated mRNA, consistent<br />

with incomplete 5 0 -end refolding <strong>of</strong> the latter<br />

species. However, DMS and T 2 probing at C120-<br />

U124 con®rmed the single-stranded loop <strong>of</strong> the target<br />

stem. In addition, V 1 cuts at C125, C126 and<br />

A127 are consistent with stacking <strong>of</strong> the CC loop<br />

dinucleotide on a stem adenine nucleotide. The<br />

super-tac mutation increased the T 2 cleavages in the<br />

loop region (C120 to U124), whereas it reduced<br />

cleavage at U117, G118 and A127-A in the stem.<br />

Thus, the super-tac mutation seems to stabilize the<br />

target hairpin, consistent with incomplete refolding<br />

<strong>of</strong> the wt truncated RNA. This conclusion was corroborated<br />

<strong>by</strong> the V 1 cuts at C139 in the dcb<br />

element. These cuts increase gradually in the order:<br />

full-length < truncated < super-tac truncated mRNA<br />

consistent with the proposed transition <strong>of</strong> the dcb/<br />

TIR <strong>hok</strong> interaction in full-length RNA to the SD mok/<br />

dcb interaction in truncated RNA.<br />

The secondary structure model <strong>of</strong> truncated <strong>hok</strong><br />

mRNA shown in Figure 2D predicts that <strong>sok</strong>T 0 and<br />

the mok TIR are located within the same hairpin.<br />

The validity <strong>of</strong> our secondary structure model <strong>of</strong><br />

the truncated mRNA was further investigated<br />

using an RNase H cleavage assay that tests the<br />

propensity <strong>of</strong> an RNA to hybridize with different<br />

DNA oligonucleotides (Zarrinkar & Williamson,<br />

1994). The DNA oligonucleotides used here are<br />

indicated in Figure 2.<br />

The cleavage pattern <strong>of</strong> full-length <strong>hok</strong> mRNA is<br />

shown in the ®rst panel <strong>of</strong> Figure 5. Oligo 1<br />

resulted in 6% cleavage only, and oligos 2 and 3<br />

both yielded negligible cleavage. This pattern is<br />

Figure 5. RNase H cleavage patterns <strong>of</strong> 32 P-labelled fulllength<br />

(®rst panel), truncated (second panel) and supertac<br />

truncated mRNA (third panel). The regions <strong>of</strong> complementarity<br />

to the oligonucleotides used are shown in<br />

Figure 2. Cleavage products are marked <strong>by</strong> arrowheads.<br />

consistent with a compact secondary structure,<br />

especially in the region <strong>of</strong> <strong>sok</strong>T 0 . In contrast, all<br />

three oligonucleotides yielded signi®cant RNase H<br />

cleavage <strong>of</strong> truncated <strong>hok</strong> mRNA (second panel). In<br />

this case, 38%, 14% and 50% cleavage was<br />

observed with oligonucleotides 1, 2 and 3, respectively.<br />

The third panel shows the cleavage pattern<br />

<strong>of</strong> super-tac truncated mRNA. Here, oligos 1, 2 and<br />

3 yielded 77%, 19% and 65% cleavage, respectively.<br />

These two latter sets <strong>of</strong> mapping data are consistent<br />

with the formation <strong>of</strong> the antisense target<br />

stem±loop presented in Figure 2D. In addition,<br />

oligo 3, which is complementary to <strong>sok</strong>T 0 , yielded<br />

more than a 100-fold difference in cleavage activity<br />

between full-length and truncated mRNA. This<br />

pattern is consistent with the difference in antisense<br />

RNA binding described above. Oligo 1,<br />

which is complementary to SD mok, showed an<br />

increased cleavage in super-tac truncated mRNA as<br />

compared to the wild-type, consistent with the<br />

observed tw<strong>of</strong>old increase in translation (Franch &<br />

Gerdes, 1996). In addition, the increased cleavage<br />

<strong>of</strong> the super-tac mRNA with all three oligonucleotides<br />

corroborates the suggestion that the<br />

wild-type truncated mRNA in vitro exists in two<br />

con®gurations, a refolded and a non-refolded. This<br />

further suggests that the antisense RNA target<br />

stem±loop favoured <strong>by</strong> the tac/ucb and SD mok/dcb<br />

pairings may be in a dynamic equilibrium with the<br />

non-refolded structure shown in Figure 2C.


44 Structure and Function <strong>of</strong> <strong>hok</strong> mRNA <strong>of</strong> <strong>Plasmid</strong> R1<br />

Figure 6. Primary and proposed local secondary structures <strong>of</strong> wild-type and mutated antisense RNA target hairpins.<br />

The SD mok and AUG mok sequences are marked <strong>by</strong> a black line and the dcb box <strong>by</strong> a hairline. Mutations are shown in<br />

bold. The energy <strong>of</strong> the local structures is given below each structure (calculated <strong>by</strong> MFOLD). The in vitro binding<br />

assays and translation data from Figures 7 and 8 are presented here in a quantitative form.<br />

The internal loop <strong>of</strong> the target hairpin is<br />

required for antisense RNA binding<br />

The loop <strong>of</strong> the target hairpin in truncated <strong>hok</strong><br />

mRNA is complementary to the 5 0 single-stranded<br />

end <strong>of</strong> Sok-RNA (Figure 2A and D). This suggests<br />

that the stem±loop structure is crucial for optimal<br />

antisense RNA binding. Therefore, we tested the<br />

function <strong>of</strong> the internal loop in antisense RNA<br />

Figure 7. In vitro binding between truncated <strong>hok</strong> mRNA<br />

containing the tst mutations and 32 P-labelled Sok-RNA<br />

shown <strong>by</strong> gel-shift. The Sok-RNAs used to bind tst2 and<br />

tst1,2 target RNAs carried cognate mutations to maintain<br />

complete antisense/target complementarity.<br />

binding. Structure-closing mutations yielding a<br />

non-interrupted upper stem were introduced at<br />

both sides <strong>of</strong> the target stem (tst1 and tst2). The<br />

double tst1,2 mutation was constructed in order to<br />

regain the internal loop. The secondary structures<br />

<strong>of</strong> the mutated target RNAs are shown in Figure 6.<br />

Antisense binding assays were accomplished as<br />

described above using antisense RNAs with 5 0 -<br />

ends completely complementary to the mutated<br />

<strong>hok</strong> target RNAs. Typical binding assays are shown<br />

in Figure 7, and the relative binding-rates calculated<br />

from three independent experiments are<br />

given in Figure 6. The observed binding rates for<br />

the truncated mRNA carrying tst1 and tst2 were<br />

reduced approximately eightfold and 16-fold,<br />

respectively. The double mutation (tst1,2) completely<br />

restored the binding rate to that <strong>of</strong> the wildtype<br />

RNAs. These results show (i) that interrupted<br />

helicity in the target stem is required for optimal<br />

antisense RNA binding and (ii) that the postulated<br />

secondary structure <strong>of</strong> the target hairpin is valid.<br />

The internal loop <strong>of</strong> the target hairpin is<br />

required for translation <strong>of</strong> <strong>hok</strong> (and mok)<br />

The effect <strong>of</strong> the tst mutations on <strong>hok</strong> expression<br />

was tested <strong>by</strong> in vitro translation <strong>of</strong> truncated<br />

mRNAs. Note, that this assay detects translation <strong>of</strong><br />

<strong>hok</strong>, not mok. As shown in Figure 8A, the tst1 and


Structure and Function <strong>of</strong> <strong>hok</strong> mRNA <strong>of</strong> <strong>Plasmid</strong> R1 45<br />

Figure 8. A and B, SDS-PAGE <strong>of</strong> in vitro translation<br />

reactions <strong>of</strong> truncated <strong>hok</strong> mRNA. Relative translation<br />

ef®ciencies are normalized to LacZ(a) and b-lactamase<br />

internal standards, and their quanti®cations are given in<br />

Figure 6.<br />

tst2 mutations abolished <strong>hok</strong> translation. The tst1,2<br />

mutation partially restored translation (to 9% <strong>of</strong><br />

that <strong>of</strong> the wild-type). The tst1,2 mutation increases<br />

the G value <strong>of</strong> the hairpin <strong>by</strong> 3.4 kcal/mol<br />

because <strong>of</strong> the G-U base-pairing partially closing<br />

the internal loop (Figure 6). Therefore, an<br />

additional mutation containing an inverted internal<br />

loop sequence was constructed and tested (tst-inv).<br />

This mutation reduced the <strong>hok</strong> translation rate to<br />

17% (Figure 8A). These results are consistent with<br />

the notion that the internal loop is responsible for a<br />

suf®ciently low hairpin stability as to allow loading<br />

<strong>of</strong> ribosomes at the mok TIR.<br />

Mutations in the downstream complementary<br />

box (dcb) reduce <strong>hok</strong> expression<br />

In the target hairpin, the SD mok element is proposed<br />

to base-pair with the downstream complementary<br />

box, dcb (Figures 2D and 6). This<br />

sequestration was puzzling, since it is known that<br />

secondary structures tend to reduce or even prevent<br />

translation (de Smit & van Duin, 1990a,b),<br />

and <strong>hok</strong> mRNA is translated at a signi®cant rate<br />

(Figure 8). To investigate this apparent discre-<br />

pancy, mutations that reduce the pairing to SD mok<br />

were introduced into the dcb element (Figure 6).<br />

Surprisingly, both dcb mutations resulted in a<br />

severe reduction <strong>of</strong> <strong>hok</strong> translation (Figure 8A).<br />

Thus, the SD mok/dcb interaction appears to be<br />

required to maintain the wild-type translation rate<br />

<strong>of</strong> <strong>hok</strong> mRNA.<br />

A straightforward explanation for this unexpected<br />

result could be that the SD mok/dcb interaction<br />

is required for 5 0 refolding and therefore for<br />

translational activation. Hence, weakening <strong>of</strong> the<br />

SD mok/dcb interaction would favour the nonrefolded<br />

ucb/SD mok con®guration, thus explaining<br />

the effect <strong>of</strong> the dcb mutations on translation. To<br />

test this inference, the super-tac mutation was<br />

combined with the dcb1 and dcb2 mutations and<br />

the respective RNAs subjected to translation<br />

(Figure 8B). The tst1,2 and tst-inv mutations were<br />

also combined with super-tac and translation <strong>of</strong> the<br />

respective RNAs included for comparison. The<br />

relative translation-rates <strong>of</strong> tst1,2 and tst-inv were<br />

not signi®cantly increased <strong>by</strong> the super-tac refolding<br />

mutation. This was expected, since the tst1,2<br />

and tst-inv mutations stabilize the target stem (see<br />

Figure 6). In contrast, the super-tac mutation<br />

resulted in a three- to sixfold increase in <strong>hok</strong> translation<br />

<strong>of</strong> RNAs carrying the dcb1 and dcb2<br />

mutations (Figure 6). Thus, the super-tac mutation<br />

could, at least partially, reverse the translational<br />

defect conferred <strong>by</strong> the dcb mutations. However,<br />

the super-tac dcb mRNAs were still translated at a<br />

rate signi®cantly lower than that <strong>of</strong> the wild-type<br />

truncated RNA (Figure 8B).<br />

Toeprinting <strong>of</strong> the mok TIR in full-length and<br />

truncated <strong>hok</strong> mRNAs<br />

To investigate the accessibility <strong>of</strong> SD mok more<br />

directly, we employed the toeprinting technique.<br />

The toeprint assay semi-quantitatively monitors<br />

the binding af®nity between a TIR and the ribosomal<br />

30 S subunit (Hartz et al., 1988). Our toeprint<br />

analysis <strong>of</strong> the mok TIR is shown in Figure 9. In<br />

truncated <strong>hok</strong> mRNA, a weak, but clearly discernible<br />

band corresponding to a reverse transcriptase<br />

stop at ‡15 relative to the mok AUG start-codon<br />

was detected (lane 4). The position <strong>of</strong> this toeprint<br />

is consistent with a ternary complex formed at the<br />

mok AUG codon. We were not able to detect a<br />

similar toeprint in full-length <strong>hok</strong> mRNA (lane 2).<br />

This correlates well with the lack <strong>of</strong> translation <strong>of</strong><br />

full-length <strong>hok</strong> mRNA in vivo and in vitro (Gerdes<br />

et al., 1990; Thisted et al., 1994a). Furthermore, the<br />

weak mok toeprint in truncated <strong>hok</strong> mRNA<br />

suggests that it is translated relatively infrequently,<br />

consistent with the pairing <strong>of</strong> the mok<br />

TIR to the dcb element in the target RNA hairpin.<br />

Similar results were obtained <strong>by</strong> the use <strong>of</strong><br />

another oligonucleotide as primer in the toeprinting<br />

assay, indicating that the quantitative aspects<br />

described here are independent <strong>of</strong> the actual<br />

experimental setup.


46 Structure and Function <strong>of</strong> <strong>hok</strong> mRNA <strong>of</strong> <strong>Plasmid</strong> R1<br />

Figure 9. In vitro toeprinting analysis <strong>of</strong> full-length and truncated <strong>hok</strong> mRNAs. The <strong>hok</strong>1 primer was used in all reactions<br />

(see Materials and Methods). Reverse transcription stops (toeprints) on wild-type and mutated <strong>hok</strong> mRNAs at<br />

AUG mok ‡ 15 are marked <strong>by</strong> squares. Lanes ‡ contain mRNA incubated with tRNA fMet and 30 S subunits. Lanes<br />

are control reactions without tRNA fMet and 30 S subunits. The type <strong>of</strong> RNA used in each toeprint is indicated for each<br />

pair <strong>of</strong> lanes. All mRNAs used are truncated <strong>hok</strong> mRNA derivatives, except for lanes 1 and 2. The tst and dcb<br />

mutations are shown in Figure 6, and super-tac in Figure 2.<br />

The accessibility <strong>of</strong> the mok TIR determines<br />

the rate <strong>of</strong> <strong>hok</strong> translation<br />

To investigate whether the translatability <strong>of</strong> <strong>hok</strong><br />

mRNA and toeprinting at the mok TIR appear to<br />

be correlated, we performed toeprinting analyses<br />

<strong>of</strong> mutated <strong>hok</strong> mRNAs. As seen from lanes 6<br />

and 8 <strong>of</strong> Figure 9, the presence <strong>of</strong> the tst1 and<br />

tst2 structure-closing mutations completely abolished<br />

toeprinting at the mok TIR, consistent with<br />

the lack <strong>of</strong> translation <strong>of</strong> these RNAs (Figure 6).<br />

The tst1,2 and tst-inv mRNAs yielded reduced<br />

toeprinting signals (lanes 10 and 12), again consistent<br />

with the degree <strong>of</strong> translation <strong>of</strong> the<br />

mRNAs. The mRNAs containing the dcb1 and<br />

dcb2 mutations yielded slightly reduced toeprint<br />

signals (lanes 14 and 16) and their translation<br />

rates were also clearly reduced (Figure 6). Furthermore,<br />

the super-tac mutation that previously<br />

was shown to enhance translation <strong>of</strong> <strong>hok</strong> <strong>by</strong> tw<strong>of</strong>old,<br />

yielded a threefold increase in the mok toeprinting<br />

signal (lane 18). These results indicate<br />

that the strength <strong>of</strong> the toeprinting signal from<br />

the mok TIR and the <strong>hok</strong> translation rate are correlated.<br />

We next asked if the combination <strong>of</strong> the<br />

super-tac mutation and the tst1,2 and tst-inv<br />

mutations would enhance the toeprinting signal<br />

(lanes 20 and 22, respectively). As seen, the<br />

super-tac mutation in these cases did not lead to<br />

an increased toeprinting signal. This is consistent<br />

with the lack <strong>of</strong> enhanced <strong>hok</strong> translation <strong>by</strong><br />

super-tac in these two cases (Figure 6). Thus, all<br />

our observations indicate that within limits, the<br />

rate <strong>of</strong> ribosome loading at the mok TIR determines<br />

the rate <strong>of</strong> <strong>hok</strong> translation.<br />

The SD mok/dcb interaction is required for<br />

loading <strong>of</strong> ribosomes at the mok TIR<br />

Even though super-tac to a signi®cant degree<br />

reverted the negative effect <strong>of</strong> the dcb mutations,<br />

the double super-tac dcb mutants were not translated<br />

as ef®ciently as the wild-type <strong>hok</strong> mRNA<br />

(Figure 6). One explanation could be that in the<br />

double super-tac dcb mutant mRNA mok translation<br />

is so intense that it reduces translation <strong>of</strong> <strong>hok</strong>.<br />

Examples <strong>of</strong> occluding overlapping translation<br />

have been described (Berkhout et al., 1985). In the<br />

latter case, toeprinting at the mok TIR yields much<br />

more direct information than the <strong>hok</strong> translationrate.<br />

As seen from Figure 9, the super-tac mutation<br />

enhanced the toeprinting signals <strong>of</strong> both dcb<br />

mutant mRNAs more than tenfold (compare lanes<br />

14 and 16 with lanes 24 and 26), and these RNAs<br />

clearly yielded the strongest signals <strong>of</strong> all the<br />

mutants. These results are most readily explained<br />

<strong>by</strong> the assumption that super-tac forced refolding <strong>of</strong><br />

the mRNA into the SD mok/dcb con®guration, combined<br />

with the reduced base-pairing between<br />

SD mok and dcb together cause the severe increase in<br />

the toeprinting signals <strong>of</strong> the doubly mutated<br />

mRNAs. Taken together, our results show that the<br />

antisense RNA hairpin is ®nely tuned to its two<br />

functions, translation and antisense RNA binding,<br />

since virtually any change in its structure has detrimental<br />

effects.<br />

Discussion<br />

Here, we obtain direct evidence for the refolding<br />

model proposed in the accompanying paper


Structure and Function <strong>of</strong> <strong>hok</strong> mRNA <strong>of</strong> <strong>Plasmid</strong> R1 47<br />

(Gultyaev et al., 1997). We predict therein that the<br />

3 0 processing <strong>of</strong> the inactive full-length <strong>hok</strong> mRNA<br />

triggers major structural rearrangements at its 5 0 -<br />

end that activate translation and antisense RNA<br />

binding. Accordingly, we show here that the two<br />

versions <strong>of</strong> <strong>hok</strong> mRNA have highly different properties<br />

with respect to translation and antisense<br />

RNA binding, and subsequently correlate the biological<br />

properties <strong>of</strong> the molecules with their foldings.<br />

Finally, we dissect the antisense RNA target<br />

hairpin and show that this structure is required for<br />

optimal antisense RNA binding and for proper<br />

translation.<br />

Full-length <strong>hok</strong> mRNA bound Sok-RNA at a negligible<br />

rate in vivo (Figure 3A) and in vitro<br />

(Figure 3B). These results are consistent with the<br />

secondary structure model <strong>of</strong> the RNA as shown<br />

in Figure 2B: the Sok-RNA target region (<strong>sok</strong>T 0 ) is<br />

base-paired with the 3 0 -end <strong>of</strong> the RNA and is<br />

therefore unable to react with the 5 0 -end <strong>of</strong> the<br />

antisense RNA. Furthermore, the fbi element<br />

located at the very 3 0 -end <strong>of</strong> the mRNA pairs with<br />

the 5 0 tac element. This pairing prevents tac stem<br />

formation and forces the ucb element to pair with<br />

SD mok. By toeprinting analyses (Figure 9), we show<br />

that this pairing prevented access <strong>of</strong> the ribosomes<br />

at SD mok, which is consistent with the lack <strong>of</strong> translation<br />

<strong>of</strong> full-length <strong>hok</strong> mRNA (Thisted et al.,<br />

1994a).<br />

In contrast, truncated <strong>hok</strong> mRNA bound Sok-<br />

RNA avidly (Figure 3) and was translated<br />

ef®ciently (Figure 6). Truncated <strong>hok</strong> mRNA is generated<br />

<strong>by</strong> slow continuous 3 0 processing <strong>of</strong> the<br />

full-length mRNA (Gerdes et al., 1990; Thisted et al.,<br />

1994a) in both plasmid-carrying and plasmid-free<br />

cells. In the former case, the presence <strong>of</strong> Sok-RNA<br />

prevents translation <strong>of</strong> truncated <strong>hok</strong> mRNA and<br />

mediates its rapid decay via RNase III-mediated<br />

hydrolysis (Figure 3A, ®rst panel). <strong>Plasmid</strong>-free<br />

cells, however, do not contain antisense RNA, and<br />

truncated <strong>hok</strong> mRNA therefore accumulates<br />

(Figure 3A, second panel). We have shown previously<br />

that the 3 0 processing leads to Hok protein<br />

synthesis and killing <strong>of</strong> the plasmid-free cells<br />

(Gerdes et al., 1990; Thisted et al., 1994a). Thus, the<br />

post-segregational killing mechanism is dependent<br />

on the formation <strong>of</strong> truncated <strong>hok</strong> mRNA.<br />

In the accompanying paper we suggest that the<br />

exonucleolytic removal <strong>of</strong> the <strong>hok</strong> mRNA 3 0 -end<br />

triggers refolding <strong>of</strong> the 5 0 -end with formation <strong>of</strong><br />

the tac stem and the antisense RNA target hairpin.<br />

Figure 2 shows the structural rearrangements proposed<br />

to activate both translation and antisense<br />

RNA binding. In truncated RNA, tac now prefers<br />

to pair with the ucb element (Figure 2D). This interaction<br />

favours the formation <strong>of</strong> the antisense RNA<br />

target hairpin. In this hairpin, SD mok pairs with the<br />

dcb element (Figure 2D). The AUG mok start codon is<br />

also paired, but an internal loop is located between<br />

the start codon and SD mok. The 5 0 -end <strong>of</strong> Sok-RNA<br />

is complementary to six <strong>of</strong> the seven nucleotides in<br />

the loop <strong>of</strong> the target hairpin. We have shown<br />

recently that these nucleotides are involved in the<br />

initial recognition reaction between the antisense<br />

RNA and the mRNA, consistent with their singlestranded<br />

conformation (unpublished results). The<br />

analysis <strong>of</strong> the antisense RNA recognition reaction<br />

will be presented elsewhere.<br />

To obtain direct evidence for the proposed structural<br />

rearrangements, we accomplished secondary<br />

structure analyses (Figure 4) and RNase H cleavage<br />

mapping (Figure 5) <strong>of</strong> the two RNAs. Data<br />

obtained from both types <strong>of</strong> experiments are largely<br />

consistent with the structure models shown in<br />

Figure 2B and D. Clearly, the introduction <strong>of</strong> the<br />

super-tac mutation (Figure 2D) into truncated <strong>hok</strong><br />

mRNA changed the cleavage pattern further<br />

downstream in the molecule (Figure 4). The altered<br />

cleavage pattern is best explained <strong>by</strong> the assumption<br />

that wild-type truncated RNA exists in two<br />

con®gurations, a refolded and a non-refolded<br />

form, and that the super-tac mutation forces refolding<br />

<strong>of</strong> the entire population <strong>of</strong> molecules. This<br />

interpretation is consistent with our translation<br />

and toeprinting data (see below).<br />

Phylogeny and genetic algorithm calculations<br />

support the suggestion that all the <strong>hok</strong>-homologous<br />

gene systems form antisense RNA target hairpins<br />

<strong>of</strong> similar sizes and energies (see the accompanying<br />

paper, Gultyaev et al., 1997). Therefore, we investigated<br />

the function <strong>of</strong> the target hairpin with<br />

respect to translation and antisense RNA binding.<br />

To this end, we introduced a number <strong>of</strong> speci®c<br />

mutations in the stem and assayed their effects<br />

(summarized in Figure 6). First, the stem-closing<br />

mutations (tst1 and tst2) completely abolished<br />

translation and severely impeded antisense RNA<br />

binding. The double tst1,2 re-opening mutation<br />

fully restored antisense RNA binding, and translation<br />

was partially regained. These results show<br />

that the internal loop is required for the proper<br />

function <strong>of</strong> the target hairpin. The tst-inv mutation,<br />

in which the nucleotides <strong>of</strong> the internal loop were<br />

inverted, also exhibited a reduced translation rate<br />

(Figure 6). The latter two mutations show that the<br />

wild-type stem is accurately tuned to its particular<br />

translation rate, since even modest changes in its<br />

stability pr<strong>of</strong>oundly in¯uenced the translation rate.<br />

In the RNAs described until now, the ( G values<br />

<strong>of</strong> the target hairpin and the translation-rates are<br />

inversely correlated with the stability <strong>of</strong> the hairpin<br />

(Figure 6).<br />

The pairing <strong>of</strong> SD mok with dcb was puzzling,<br />

since sequestration <strong>of</strong> SD elements is known to<br />

reduce or prevent translation (de Smit & van Duin,<br />

1990a,b, 1994). However, phylogeny and structureprobing<br />

data strongly support the postulated interaction.<br />

Point mutations in dcb that reduced the<br />

SD mok/dcb interaction therefore should lead to an<br />

increased translation rate. However, the dcb1 and<br />

dcb2 mutations both exhibited severely decreased<br />

rates <strong>of</strong> <strong>hok</strong> translation (eightfold and 16-fold,<br />

respectively; see Figure 6). To some extent this<br />

result was surprising and indicated that the point<br />

mutations perhaps shift the mok TIR into an<br />

alternative, non-translatable con®guration. One


48 Structure and Function <strong>of</strong> <strong>hok</strong> mRNA <strong>of</strong> <strong>Plasmid</strong> R1<br />

straight-forward possibility would be that the dcb<br />

mutations favour the ucb/SD mok interaction present<br />

in the full-length RNA. As shown <strong>by</strong> toeprinting,<br />

this RNA con®guration prevents access <strong>of</strong> ribosomes<br />

at SD mok (Figure 9). To test this assumption,<br />

we combined the super-tac mutation with the dcb<br />

mutations, since super-tac predictably would shift<br />

the equilibrium towards the SD mok/dcb interaction<br />

and there<strong>by</strong> stimulate ribosome-binding at the mok<br />

TIR and thus, translation <strong>of</strong> <strong>hok</strong>. The super-tac<br />

mutation was also combined with the tst1,2, and<br />

tst-inv mutations. The super-tac mutation stimulated<br />

<strong>hok</strong> translation <strong>of</strong> the wild-type mRNA tw<strong>of</strong>old,<br />

consistent with the assumption that super-tac<br />

forces refolding <strong>of</strong> the entire population <strong>of</strong> molecules<br />

(Figure 6). In contrast, super-tac did not<br />

stimulate translation <strong>of</strong> RNAs containing tst1,2 and<br />

tst-inv. The latter result shows that translation <strong>of</strong><br />

these RNAs is refractory to stimulation <strong>by</strong> supertac.<br />

The target hairpins containing tst1,2 and tst-inv<br />

are more energy-rich than the wt hairpin (Figure 6).<br />

Thus, the lack <strong>of</strong> stimulation <strong>by</strong> super-tac suggests<br />

that these RNAs are already fully shifted towards<br />

the refolded, active con®guration.<br />

In the cases <strong>of</strong> dcb1 and dcb2, super-tac stimulated<br />

translation threefold and sixfold, respectively (i.e.<br />

the stimulation was signi®cant). This result is consistent<br />

with the proposed structural change conferred<br />

<strong>by</strong> the dcb mutations. However, even<br />

though super-tac stimulated <strong>hok</strong> translation <strong>of</strong> the<br />

dcb mutant RNAs, the translation rates <strong>of</strong> the doubly<br />

mutated RNAs never exceeded that <strong>of</strong> the<br />

wild-type RNA. This was surprising, since supertac<br />

forces formation <strong>of</strong> the translatable con-<br />

®guration and the dcb mutations reduce the<br />

sequestration <strong>of</strong> SD mok. However, these observations<br />

are indirect, since we change the environment<br />

<strong>of</strong> the mok TIR and measure <strong>hok</strong> translation.<br />

This is, <strong>of</strong> course, a technical caveat.<br />

To gain more direct insight into the status <strong>of</strong> the<br />

mok TIR, we turned to the toeprinting technique<br />

(Figure 9). Most importantly, we found that in all<br />

but two cases there was a correlation between the<br />

in vitro translation rates <strong>of</strong> <strong>hok</strong> and the strength <strong>of</strong><br />

the corresponding SD mok toeprinting signal. In the<br />

cases exempt, dcb1 and dcb2, respectively, super-tac<br />

severely enhanced the toeprinting signals relative<br />

to the translation rates. These results show that the<br />

dcb mutations lead, as expected, to a highly accessible<br />

SD mok element when combined with super-tac.<br />

The relatively slow translation rates from the<br />

double super-tac dcb mutant RNAs may be caused<br />

<strong>by</strong> the highly increased mok translation rate, which<br />

inhibits <strong>hok</strong> translation (Berkhout et al., 1985). This<br />

inference is consistent with the presence <strong>of</strong> unexpected<br />

Mok protein bands appearing in the translation<br />

reactions <strong>of</strong> these RNAs (data not shown).<br />

The results discussed above suggest that the<br />

SD mok/dcb interaction prevents the RNA from folding<br />

into the non-translatable con®guration. Only<br />

when the non-translatable con®guration is disfavoured<br />

<strong>by</strong> other structural mutations (i.e. supertac),<br />

the dcb element appears, as expected, to have<br />

a negative effect on translation. These assumptions<br />

are supported <strong>by</strong> foldings <strong>of</strong> the mutated RNAs<br />

using the genetic algorithm (not shown). Thus, the<br />

toeprinting data explain the somewhat surprising<br />

observation that an anti-SD element (dcb) is<br />

required to maintain an SD element in a translatable<br />

con®guration. Our results are still compatible<br />

with and actually support the paradigm that secondary<br />

structure interactions at SD elements certainly<br />

reduce translation (de Smit & van Duin,<br />

1990a, 1994).<br />

In bacteria, exonucleolytic 3 0 processing is a<br />

major pathway <strong>of</strong> mRNA inactivation and decay<br />

(Higgins et al., 1988; Belasco & Higgins, 1988). We<br />

believe that the <strong>hok</strong> and <strong>hok</strong>-homologous mRNAs<br />

represent a unique class <strong>of</strong> mRNAs that are activated<br />

<strong>by</strong> 3 0 processing and we know <strong>of</strong> no other<br />

example in which an RNA is activated <strong>by</strong> 5 0 refolding<br />

triggered <strong>by</strong> the removal <strong>of</strong> its 3 0 -end. However,<br />

several other mRNAs contain alternative<br />

competing structures involved in translational control.<br />

The rpsO mRNA encoding the ribosomal protein<br />

S15 and the exoribonuclease PNPase is<br />

autoregulated <strong>by</strong> S15 that binds to the rpsO TIR.<br />

The rpsO mRNA leader region exists as a translational-competent<br />

pseudoknot structure in equilibrium<br />

with a double hairpin structure occluding<br />

SD rpsO (Phillipe et al., 1990). Interestingly, the S15<br />

protein binds and sequesters the pseudoknot structure<br />

and there<strong>by</strong> inhibits initiation <strong>of</strong> translation<br />

(Phillipe et al., 1993). Also, the leader region <strong>of</strong> cIII<br />

mRNA from phage l adopts two alternative secondary<br />

structures with almost equal energies but<br />

different translational capacities (Altuvia et al.,<br />

1989). Here, the RNase III protein is believed to<br />

trap, without cleavage, the active mRNA con®guration<br />

(Altuvia et al., 1991).<br />

Many antisense RNAs contain upper-stem mismatches<br />

in order to accommodate stable antisense<br />

RNA/target interaction (Hjalt & Wagner, 1995;<br />

Kittle et al., 1989; Wilson et al., 1997). The structure<br />

<strong>of</strong> the regulatory hairpin in <strong>hok</strong> mRNA contains a<br />

four nucleotide internal loop separated from the<br />

seven nucleotide loop <strong>by</strong> four base-pairs. The<br />

stem-closing mutations tst1 and tst2 severely<br />

impaired Sok-RNA binding. A similar effect was<br />

observed for stem-closing mutations in both CopA<br />

<strong>of</strong> the R1 plasmid replication system (Hjalt &<br />

Wagner, 1995) and the RNA-OUT <strong>of</strong> Tn10 (Kittle<br />

et al., 1989). In the <strong>hok</strong>/<strong>sok</strong> system, restoring the<br />

internal loop <strong>by</strong> the tst1,2 mutation regained binding<br />

kinetics, thus indicating the requirement <strong>of</strong><br />

helix imperfection for ef®cient Sok-RNA/target<br />

pairing in vitro. Most likely, the initial binding <strong>of</strong><br />

the 5 0 -leader <strong>of</strong> Sok-RNA to the complementary six<br />

nucleotides in the target loop creates the ®rst<br />

unstable binding intermediate. Our results suggest<br />

that subsequent progression to more complete<br />

duplex formation requires interrupted helicity to<br />

favour intrastrand opening.<br />

In conclusion, we show here that the processing<br />

at the <strong>hok</strong> mRNA 3 0 -end confers dynamic<br />

rearrangements at its 5 0 -end that mediate activation


Structure and Function <strong>of</strong> <strong>hok</strong> mRNA <strong>of</strong> <strong>Plasmid</strong> R1 49<br />

<strong>of</strong> translation and antisense binding. Phylogeny<br />

and genetic algorithm calculations indicate that the<br />

other <strong>hok</strong>-homologous mRNAs exhibit similar complex<br />

rearrangements (Gultyaev et al., 1997). We<br />

believe that this is a unique example <strong>of</strong> a group <strong>of</strong><br />

phylogenetically related mRNAs that are activated<br />

<strong>by</strong> structural rearrangements triggered <strong>by</strong> 3 0 processing.<br />

Materials and Methods<br />

Enzymes and chemicals<br />

Antibiotics and chemicals was added at the following<br />

concentrations: ampicillin, 100 mg/ml; IPTG, 1 mM. All<br />

enzymes were purchased from Boehringer Mannheim<br />

unless stated otherwise.<br />

Bacterial strains<br />

The Escherichia coli K-12 strain CSH50 [ (lac pro) rpsL]<br />

(Miller, 1972) was used in the Northern transfer analysis.<br />

<strong>Plasmid</strong>s<br />

The plasmids used and constructed are as follows:<br />

plasmids pKG1540 and pKG1541 are pGEM3 (Promega)<br />

derivatives encoding the bla and lacI q genes. The plasmids<br />

carry a mutant 450 bp <strong>hok</strong>/<strong>sok</strong> system cloned into<br />

the EcoRI-BamHI sites <strong>of</strong> the polylinker. <strong>Plasmid</strong><br />

pKG1540 is constructed <strong>by</strong> PCR on a template carrying a<br />

stop codon in <strong>hok</strong> (Thisted et al., 1992) using the oligos (i)<br />

and (ii). Oligo (i) is 5 0 -CGGATCCAAAAAGAGTG-<br />

TTGACTTGTGAGCGGATAACAATGATACTTAGAT-T-<br />

CGGCGCTTGAGGCTTTCTGCCTCATG, italics mark<br />

the BamHI restriction site and the underlined region is<br />

complementary to ‡131 to ‡ 156 in the <strong>hok</strong>/<strong>sok</strong> locus.<br />

This oligo contains the sequence for the strongly LacI<br />

repressed P A1/04 promoter. Oligo (ii) is 5 0 -CCCCGAATTC<br />

ACAACATCAGCAAGGAGAAA, italics mark the EcoRI<br />

restriction site and the underlined region is complementary<br />

to ‡580 to ‡561 in the 3 0 -end <strong>of</strong> the <strong>hok</strong>/<strong>sok</strong> system.<br />

<strong>Plasmid</strong> pKG1541 was constructed accordingly <strong>by</strong> PCR<br />

with the same oligos on a template containing a double<br />

mutation in the 10 sequence <strong>of</strong> the Sok promoter in<br />

addition to the <strong>hok</strong> mutation described above (Thisted &<br />

Gerdes, 1992).<br />

Total RNA preparation for Northern transfer analysis<br />

Preparation <strong>of</strong> total RNA from E. coli and Northern<br />

analysis were performed as described (Gerdes et al.,<br />

1990). A culture carrying a high copy number plasmid<br />

(pUC derivative; Yanisch-Perron et al., 1985) containing<br />

the P A1/04::<strong>hok</strong> fusion and either the <strong>sok</strong> ‡ or <strong>sok</strong> genotype<br />

was grown exponentially at 37 C. At A 450 0.2,<br />

1 mM IPTG was added to induce <strong>hok</strong> transcription. After<br />

one hour the cells were pelleted and resuspended in prewarmed<br />

medium without IPTG. Samples were collected<br />

at the time-points indicated. In order to determine the<br />

precise half-life for the <strong>hok</strong> mRNAs, the dilution <strong>of</strong><br />

mRNA after IPTG withdrawal (<strong>by</strong> cell growth) was compensated<br />

for <strong>by</strong> increasing the amount <strong>of</strong> total RNA<br />

loaded onto the gel as determined from the A 450<br />

measurements. The 32 P-labelled RNA probe used was<br />

generated <strong>by</strong> phage T7 RNA polymerase and the pGEMbased<br />

plasmid pGEM342 (<strong>hok</strong> probe; Thisted & Gerdes,<br />

1992). Calculation <strong>of</strong> mRNA half-lives was accomplished<br />

<strong>by</strong> evaluation <strong>of</strong> band intensity using an LKB Ultroscan<br />

laser densitometer and the GelScan XL s<strong>of</strong>tware package<br />

(version 2.1) provided <strong>by</strong> Pharmacia.<br />

Site-directed mutagenesis<br />

The mutation shown in Figure 6 was introduced in<br />

the <strong>hok</strong>/<strong>sok</strong> system <strong>by</strong> double PCR as described <strong>by</strong><br />

Nielsen & Gerdes (1995). PCR was performed using the<br />

pBR322 vector carrying the 580 bp wild-type <strong>hok</strong>/<strong>sok</strong> system,<br />

inserted in the EcoRI - BamHI sites, as template and<br />

the pBR322 EcoRI CW and pBR322 BamHI CCW external<br />

primers. Mutant primers are: tst1, 5 0 -GGACTAGACA-<br />

TAGGGGATGC (TF17) and 5 0 -GAGGCATCCCTATG-<br />

TCTAGTCCACATAGGGATAGCCTCTTACC (TF17b);<br />

tst1,2, 5 0 -ATCCTGATGTGGACTAGACATCAGGATG-<br />

CCTC (TF17-5) and 5 0 -GAGGCATCCTGATGTCTAGTC-<br />

CACATAGGGATAGCCTCTTACC, tst-inv, 5 0 - GGCA-<br />

TCCACATGTCTAGTCCACATTCGGATAGCCTCTTA-<br />

CC (TF20a) and 5 0 -GCTATCCGAATGTGGACTAGA-<br />

CATGTGGATGCCTCGTGGTGG (TF20b); dcb1, 5 0 -<br />

CACATCAGGATAGGCTCTTACCGCGC-3 0 (TF25a) and<br />

5 0 -GCGCGGTAAGAGCCTATCCTGATGTG-3 0 (TF24b);<br />

dcb2, 5 0 - CACATCAGGATAGCGTCTTACCGCGC<br />

(TF25a) and 5 0 -GCGCGGTAAGACGCTATCCTGATGTG<br />

(TF25b). The tst2 mutation was introduced <strong>by</strong> double<br />

template PCR as described (Franch & Gerdes, 1996)<br />

using the primer 5 0 -ATCCCTATGTGGACTAGACATA<br />

(TF17-4). Mutant <strong>hok</strong>/<strong>sok</strong> fragments were all cloned in the<br />

EcoRI-BamHI restriction-sites in pUC19 or mini-R1<br />

derivatives. The plasmids carrying a mutant <strong>hok</strong>/<strong>sok</strong> system<br />

were used for PCR to generate templates for in vitro<br />

synthesis <strong>of</strong> 32 P and 3 H-labelled truncated mRNA. These<br />

transcripts were used for in vitro experiments.<br />

Synthesis <strong>of</strong> <strong>hok</strong> mRNAs in vitro<br />

Large-scale wild-type and mutant in vitro <strong>hok</strong> transcripts<br />

with different lengths were synthesized using T7<br />

RNA polymerase and DNA templates generated <strong>by</strong> PCR<br />

according to Franch & Gerdes (1996).<br />

Secondary structure probing using DMS, T 2 and V 1<br />

DMS, T 2 and V 1 structure probings were conducted<br />

according to Thisted et al. (1995), with the following<br />

exceptions. The buffer used for DMS modi®cation was<br />

as follows; 20 mM Hepes (pH 7.8), 100 mM NH 4Cl,<br />

10 mM magnesium acetate, 1 mM DTT and 10% (v/v)<br />

glycerol. RNase T 2 and V 1 cleavage reactions were carried<br />

out in TMK-glutamate buffer (20 mM Tris-acetate<br />

(pH 7.5), 10 mM magnesium acetate, 200 mM potassium<br />

glutamate).<br />

Secondary structure probing <strong>by</strong> RNase H cleavage<br />

Uniformly 32 P-labelled full-length or truncated <strong>hok</strong><br />

mRNA (20 fmol) was incubated in TMK-glutamate buffer<br />

(see above) supplemented with 1 M DTT, for three<br />

minutes at 37 C to equilibrate RNA folding. DNA oligo<br />

(10 pmol) and one unit <strong>of</strong> RNase H (Gibco) were added<br />

to give a total reaction volume <strong>of</strong> 10 ml. The reaction was<br />

incubated at 30 C for an additional 30 minutes and<br />

stopped <strong>by</strong> addition <strong>of</strong> Formamid Dye (FD) and withdrawal<br />

to 0 C. In the control reaction the oligo was<br />

omitted. Samples were heated at 80 C for three minutes<br />

prior to separation on a 5.5% polyacrylamide gel con-


50 Structure and Function <strong>of</strong> <strong>hok</strong> mRNA <strong>of</strong> <strong>Plasmid</strong> R1<br />

taining 8 M urea. The oligonucleotides used were: oligo<br />

1 (SD), 5 0 -CCTCGTGGTG; oligo 2 (stem), 5 0 -CATAGG-<br />

GATG; oligo 3 (loop), 5 0 -GTGGACTAGAC.<br />

in vitro binding reactions between <strong>hok</strong> mRNA<br />

and Sok-RNA<br />

All binding reactions were performed according<br />

to Thisted et al. (1994b), except that the <strong>hok</strong> mRNA<br />

and Sok-RNA concentrations were 3 10 8 M and<br />

3 10 9 M, respectively. Also, the buffer used was TMK<br />

(see above). Band intensity quanti®cation was done <strong>by</strong><br />

Phosphor Imager and the Imagequant s<strong>of</strong>tware package<br />

(Molecular Dynamics).<br />

In vitro translation<br />

The E. coli coupled transcription/translation system<br />

(Zubay, 1973) was purchased from Promega. The translation<br />

reactions contained the following in a total volume<br />

<strong>of</strong> 20 ml: 3 pmol <strong>of</strong> 3 H-labelled truncated <strong>hok</strong> mRNA in<br />

5 ml <strong>of</strong> TE, 1 mg <strong>of</strong> pUC19 plasmid (internal standard),<br />

0.5 mM each <strong>of</strong> the 20 amino acids minus methionine,<br />

0.2 mM L-[ 35 S]methionine (1000 Ci/mmol; NEN), 2 mM<br />

ATP, 0.5 mM each <strong>of</strong> GTP, UTP and CTP, 210 mM potassium<br />

glutamate, 20 mM phosphoenolpyruvate, 35 mM<br />

Tris-acetate (pH 8.0), 27 mM ammonium acetate, 1 mM<br />

cAMP, 20 g/ml folinic acid, 100 mg/ml E. coli tRNAs,<br />

9 mM magnesium acetate, 0.8 mM IPTG, 2 mM DTT,<br />

35 mg/ml PEG 8000 and 6 ml <strong>of</strong> E. coli S30 extract. The<br />

reactions were allowed to proceed for 30 minutes. The<br />

samples were precipitated with four volumes <strong>of</strong> acetone<br />

at 0 C, dried and redissolved in 2 SDS sample buffer<br />

(per 10 ml: 2 ml <strong>of</strong> glycerol, 2 ml <strong>of</strong> 10% (w/v) SDS,<br />

0.25 mg <strong>of</strong> bromophenol blue, 2.5 ml <strong>of</strong> 0.5 mM Tris-HCl<br />

(pH 6.8), 0.4% SDS, 0.5 ml <strong>of</strong> b-mercaptoethanol) and<br />

denatured at 100 C for ®ve minutes prior to loading<br />

onto SDS/polyacrylamide gels according to Thisted et al.<br />

(1994a).<br />

Toeprinting analyses<br />

The 30 S ribosomal subunits were puri®ed as follows:<br />

100 ml <strong>of</strong> mid-log culture <strong>of</strong> the E. coli strain MRE600<br />

(RNase I-de®cient) was snap-cooled and cells were harvested.<br />

The cell pellet was resuspended in 1 ml <strong>of</strong> ribosome<br />

buffer (50 mM Tris-HCl (pH 7.5), 1 mM MgCl 2,<br />

100 mM NH 4Cl). <strong>Cell</strong> membrane fracturing was accomplished<br />

<strong>by</strong> three cycles <strong>of</strong> snap-freezing and thawing (on<br />

ice). The suspension was pelleted and the supernatant<br />

containing uncoupled ribosome subunits was ultracentrifuged<br />

(18,000 rpm, SW28 buckets) in a 5% to 20% (w/v)<br />

sucrose gradient in ribosome buffer at 4 C for 16 hours.<br />

The 30 S ribosome fractions were collected and pelleted<br />

<strong>by</strong> ultracentrifugation. The 30 S ribosome subunits were<br />

dissolved in ribosome buffer containing 10 mM MgCl 2.<br />

The toeprinting analysis was conducted as follows:<br />

0.4 pmol <strong>of</strong> 5 0 -end labelled <strong>hok</strong>1 primer was annealed to<br />

0.4 pmol <strong>of</strong> <strong>hok</strong> mRNA in a buffer containing 60 mM<br />

NH 4Cl, 10 mM Tris-acetate (pH 7.5), 8.5 mM b-mercaptoethanol,<br />

10 mM MgCl 2 supplemented with ten units <strong>of</strong><br />

RNAguard (Pharmacia) and 100 mM <strong>of</strong> dNTP and preincubated<br />

at 37 C for ®ve minutes. The 30 S ribosomes<br />

(2 ml <strong>of</strong> 0.2 mM) were added followed <strong>by</strong> ten minutes<br />

incubation. Addition <strong>of</strong> 1 ml (10 mM) <strong>of</strong> uncharged<br />

tRNA fMet (Sigma) was succeeded <strong>by</strong> 15 minutes incubation<br />

before supplementing with 1 ml (two units) <strong>of</strong><br />

AMV-RT. Reactions were stopped after 20 minutes <strong>by</strong><br />

precipitation in ethanol and subsequently resuspended<br />

in 5 ml <strong>of</strong> FD. An RNA sequence reaction was made<br />

according to Thisted et al. (1995). Toeprint ef®ciency was<br />

determined <strong>by</strong> Phosphor Imager and the Image-quant<br />

s<strong>of</strong>tware package (Molecular Dynamics).<br />

Acknowledgements<br />

This work was supported <strong>by</strong> the Center for Interaction,<br />

Structure, Function, and Engineering <strong>of</strong> Macromolecules<br />

(CISFEM) and the <strong>Plasmid</strong> Foundation.<br />

References<br />

Altuvia, S., Kornitzer, D., Teff, D. & Oppenheim, A. B.<br />

(1989). Alternative mRNA structures <strong>of</strong> the cIII<br />

gene <strong>of</strong> bacteriophage determine the rate <strong>of</strong> its<br />

translation initiation. J. Mol. Biol. 210, 265±280.<br />

Altuvia, S., Kornitzer, D., Kobi, S. & Oppenheim, A. B.<br />

(1991). Functional and structural elements <strong>of</strong> the<br />

mRNA <strong>of</strong> the cIII gene <strong>of</strong> the bacteriphage lambda.<br />

J. Mol. Biol. 218, 723±733.<br />

Belasco, J. G. & Higgins, C. F. (1988). Mechanisms <strong>of</strong><br />

mRNA decay in bacteria: a perspective. Gene, 72,<br />

15±23.<br />

Berkhout, B., Kastelein, R. A. & van Duin, J. (1985).<br />

Translational interference at overlapping reading<br />

frames in prokaryotic messenger RNA. Gene, 37,<br />

171±179.<br />

Case, C. C., Roels, S. M., Jensen, P. D., Lee, J., Kleckner,<br />

N. & Simons, R. W. (1989). The unusual stability <strong>of</strong><br />

the IS10 anti-sense RNA is critical for its function<br />

and is determined <strong>by</strong> the structure <strong>of</strong> its stemdomain.<br />

EMBO J. 8, 4297±4305.<br />

de Smit, M. & van Duin, J. (1990a). Secondary structure<br />

<strong>of</strong> the ribosome binding site determines the translational<br />

ef®ciency: a quantitative analysis. Proc. Natl<br />

Acad. Sci. USA, 87, 7668±7672.<br />

de Smit, M. & van Duin, J. (1990b). Control <strong>of</strong> procaryotic<br />

translation <strong>by</strong> mRNA secondary structure. Prog.<br />

Nucl. Acid Res. Mol. Biol. 38, 1±35.<br />

de Smit, M. & van Duin, J. (1994). Translational<br />

initiation on structured messengers another role for<br />

the Shine-Dalgarno interaction. J. Mol. Biol. 135,<br />

173±184.<br />

Franch, T. & Gerdes, K. (1996). <strong>Programmed</strong> cell death<br />

in bacteria: translational repression <strong>by</strong> mRNA endpairing).<br />

Mol. Microbiol. 21, 1049±1060.<br />

Gerdes, K., Helin, K., Christensen, O. W. & Lùbner-<br />

Olesen, A. (1988). Translational control and differential<br />

RNA decay are key elements regulating postsegregational<br />

expression <strong>of</strong> the killer protein<br />

encoded <strong>by</strong> the parB locus <strong>of</strong> plasmid R1. J. Mol.<br />

Biol. 203, 119±129.<br />

Gerdes, K., Thisted, T. & Martinussen, J. (1990). Mechanism<br />

<strong>of</strong> post-segregational killing <strong>by</strong> the <strong>hok</strong>/<strong>sok</strong> system<br />

<strong>of</strong> plasmid R1: the <strong>sok</strong> antisense RNA regulates<br />

the formation <strong>of</strong> a <strong>hok</strong> mRNA species correlated<br />

with killing <strong>of</strong> plasmid free cells. Mol. Microbiol. 4,<br />

1807±1818.<br />

Gerdes, K., Nielsen, A. K., Thorsted, P. & Wagner,<br />

E. G. H. (1992). Mechanism <strong>of</strong> killer gene reactivation:<br />

antisense RNA mediated RNaseIII cleavage<br />

ensures rapid turn-over <strong>of</strong> the <strong>hok</strong>, srnB and pndA<br />

effector mRNAs. J. Mol. Biol. 226, 637±649.


Structure and Function <strong>of</strong> <strong>hok</strong> mRNA <strong>of</strong> <strong>Plasmid</strong> R1 51<br />

Gultyaev, A. P., Franch, T. & Gerdes, K. (1997). <strong>Programmed</strong><br />

cell death <strong>by</strong> <strong>hok</strong>/<strong>sok</strong> <strong>of</strong> plasmid R1:<br />

coupled nucleotide covariations reveal phylogenetically<br />

conserved folding pathway in the <strong>hok</strong> family<br />

<strong>of</strong> mRNAs. J. Mol. Biol. 273, 26±37.<br />

Hartz, D., McPheeters, D. S., Traut, R. & Gold, L. (1988).<br />

Extension inhibition analysis <strong>of</strong> translation initiation<br />

complexes. Methods Enzymol. 164, 419±425.<br />

Higgins., C. F., McLaren, R. S. & Newbury, S. F. (1988).<br />

Repetitive extragenic palindromic sequences,<br />

mRNA stability and gene expression: evolution <strong>by</strong><br />

gene conversion? ± a review. Gene, 72, 3±14.<br />

Hjalt, T. AÊ . H. & Wagner, E. G. H. (1995). Bulged out<br />

nucleotides in an antisense RNA are required for<br />

target RNA binding in vitro and inhibition in vivo.<br />

Nucl. Acids Res. 23, 580±587.<br />

Kittle, J. K., Simons, R. W., Lee, J. & Kleckner, N. (1989).<br />

Insertion sequence IS10 anti-sense pairing initiates<br />

<strong>by</strong> an interaction between the 5 0 -end <strong>of</strong> the target<br />

RNA and a loop in the anti-sense RNA. J. Mol. Biol.<br />

210, 561±572.<br />

Lanzer, M. & Bujard, H. (1988). Promoters largely determine<br />

the the ef®ciency <strong>of</strong> repressor action. Proc.<br />

Natl Acad. Sci. USA, 85, 8973±8977.<br />

Miller, J. (1972). Experiments in Molecular Genetics, Cold<br />

Spring Harbor Laboratory Press, Cold Spring Harbor,<br />

NY.<br />

Nielsen, A. K. & Gerdes, K. (1995). Mechanism <strong>of</strong> postsegregational<br />

killing <strong>by</strong> <strong>hok</strong>-homologue pnd <strong>of</strong> plasmid<br />

R483: two translational control elements in the<br />

pnd mRNA. J. Mol. Biol. 249, 270±282.<br />

Phillipe, C., Portier, C., Mougel, M., Grunberg-Manago,<br />

M., Ebel, J. P., Ehresmann, B. & Ehresmann, C.<br />

(1990). Target site <strong>of</strong> Escherichia coli ribosomal protein<br />

S15 on its messenger RNA. J. Mol. Biol. 211,<br />

415±426.<br />

Phillipe, C., Eyermann, F., Benard, L., Portier, C.,<br />

Ehresmann, B. & Ehresmann, C. (1993). Ribosomal<br />

protein S15 from Esherichia coli modulates its own<br />

translation <strong>by</strong> trapping the ribosome on the mRNA<br />

initiation loading site. Proc. Natl Acad. Sci. USA, 90,<br />

4394±4398.<br />

Saenger, W. (1984). Forces stabilizing associations<br />

between bases: Hydrogen bonding and base stacking.<br />

In Principles <strong>of</strong> Nucleic Acid Structure, pp. 116±<br />

158, Springer-Verlag, New York.<br />

Siemering, K. R., Prazkier, J. & Pittard, A. J. (1994).<br />

Mechanism <strong>of</strong> binding <strong>of</strong> the antisense and target<br />

RNAs involved in the refulation <strong>of</strong> IncB plasmid<br />

replication. J. Bacteriol. 176, 2677±2688.<br />

Thisted, T. & Gerdes, K. (1992). Mechanism <strong>of</strong> post-segregational<br />

killing<strong>by</strong> the <strong>hok</strong>/<strong>sok</strong> system <strong>of</strong> plasmid<br />

R1. Sok antisense RNAregulates <strong>hok</strong> expression<br />

indirectly through the overlapping mok gene. J. Mol.<br />

Biol. 223, 41±54.<br />

Thisted, T., Nielsen, A. K. & Gerdes, K. (1994a). Mechanism<br />

<strong>of</strong> post-segregational killing: translation <strong>of</strong><br />

Hok, SrnB and Pnd mRNAs <strong>of</strong> plasmids R1, F and<br />

R483 is activated <strong>by</strong> 3 0 -end processing. EMBO J. 13,<br />

1950±1959.<br />

Thisted, T., Sùrensen, N. S., Wagner, G. H. & Gerdes, K.<br />

(1994b). Mechanism <strong>of</strong> post-segregational killing:<br />

Sok antisense RNA interacts with Hok mRNA via<br />

its 5 0 -end single-stranded leader and competes with<br />

the 3 0 -end <strong>of</strong> Hok mRNA for binding to the mok<br />

translational initiation region. EMBO J. 13, 1960±<br />

1968.<br />

Thisted, T., Sùrensen, N. S. & Gerdes, K. (1995). Mechanism<br />

<strong>of</strong> post-segregational killing: secondary structure<br />

analysis <strong>of</strong> the entire <strong>hok</strong> mRNA from plasmid<br />

R1 suggests a fold-back structure that prevents<br />

translation and antisense RNA binding. J. Mol. Biol.<br />

247, 859±873.<br />

Turner, D. H. & Bevilacqua, P. C. (1993). Thermodynamic<br />

considerations for evolution <strong>by</strong> RNA. In The<br />

RNA World (Gesteland, R. F. & Atkins, J. F., eds),<br />

pp. 447±464. Cold Spring Harbor Laboratory Press,<br />

Cold Spring Harbor, NY.<br />

Wilson, I. W., Siemering, K. R., Prazkier, J. & Pittard,<br />

A. J. (1997). Importance <strong>of</strong> structural differences<br />

between complementary RNA molecules to control<br />

<strong>of</strong> replication <strong>of</strong> an IncB plasmid. J. Bacteriol. 179,<br />

742±753.<br />

Yanisch-Perron, C., Vieira, J. & Messing, J. (1985).<br />

Improved M13 phage cloning vectors and host<br />

strains: nucleotide sequences <strong>of</strong> the M13mp18 and<br />

pUC19 vectors. Gene, 33, 103±109.<br />

Zarrinkar, P. P. & Williamson, J. R. (1994). Kinetic intermediates<br />

in RNA folding. Science, 265, 918±924.<br />

Zubay, G. (1973). In vitro synthesis <strong>of</strong> protein in<br />

microbial systems. Annu. Rev. Genet. 7, 267±287.<br />

Edited <strong>by</strong> D. E. Draper<br />

(Received 14 April 1997; received in revised form 14 July 1997; accepted 17 July 1997)

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!