Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2012 Oct 9.
Published in final edited form as: Nat Struct Mol Biol. 2008 Sep 21;15(10):1076–1083. doi: 10.1038/nsmb.1494

Crystal structures of the SAM-III/SMK riboswitch reveal the SAM-dependent translation inhibition mechanism

Changrui Lu 1, Angela M Smith 2, Ryan T Fuchs 2, Fang Ding 1, Kanagalaghatta Rajashankar 3, Tina M Henkin 2, Ailong Ke 1
PMCID: PMC3467307  NIHMSID: NIHMS254138  PMID: 18806797

Abstract

Three distinct classes of S-adenosyl-l-methionine (SAM)-responsive riboswitches have been identified that regulate bacterial gene expression at the levels of transcription attenuation or translation inhibition. The SMK box (SAM-III) translational riboswitch has been identified in the SAM synthetase gene in members of the Lactobacillales. Here we report the 2.2-Å crystal structure of the Enterococcus faecalis SMK box riboswitch. The Y-shaped riboswitch organizes its conserved nucleotides around a three-way junction for SAM recognition. The Shine-Dalgarno sequence, which is sequestered by base-pairing with the anti–Shine-Dalgarno sequence in response to SAM binding, also directly participates in SAM recognition. The riboswitch makes extensive interactions with the adenosine and sulfonium moieties of SAM but does not appear to recognize the tail of the methionine moiety. We captured a structural snapshot of the SMK box riboswitch sampling the near-cognate ligand S-adenosyl-l-homocysteine (SAH) in which SAH was found to adopt an alternative conformation and fails to make several key interactions.


Recent discoveries of cis-acting regulatory RNAs termed riboswitches revealed the novel ability of RNA to directly relay environmental cues to the genetic regulation machinery (for reviews, see refs. 13 and references therein). Bacteria use riboswitches to regulate the metabolism and transport of vitamins49, nucleotides1015, amino acids1619, cofactors5,2024 and metal ions25,26. Riboswitches are especially prevalent in Gram-positive bacteria, as exemplified by Bacillus subtilis, where more than 4% of its genes are riboswitch regulated27. A handful of riboswitches were also found in eukaryotic systems, suggesting that such mechanisms may be even more widespread than currently appreciated9,28,29.

A typical metabolite binding riboswitch consists of a ligand-sensing ‘aptamer domain’ and an ‘output domain’ that regulates gene expression, generally at the levels of transcription attenuation or translation initiation2,4,30. Binding of the cognate metabolite to the aptamer domain usually causes conformational changes that result in alternative base-pairing in the output domain, which in turn affects expression of the downstream open reading frames. The three classes of SAM-responsive riboswitches, including the S box (SAM-I)2022,31, SAM-II23 and SMK box (SAM-III)24, represent the most commonly found riboswitch families in bacterial genomes32. All of these riboswitches are highly selective for SAM over its natural analog SAH, which is generated from utilization of SAM as a methyl donor in enzyme-catalyzed methyltransfer reactions and differs from SAM by the absence of a methyl group and a positive charge on the sulfur atom. Crystal structures of the S box (SAM-I) and SAM-II riboswitches33,34 revealed two completely different RNA folds that use distinct mechanisms of SAM recognition.

The SMK box riboswitch (Fig. 1a) regulates the translation of SAM synthetase (metK) genes in lactic acid bacteria by sequestration of the Shine-Dalgarno (SD) sequence, which is essential for loading of the 30S ribosomal subunit for translation initiation24. Notably, the SD sequence within the SMK box family (GGGGG) differs from the consensus SD sequence (GGAGG) at the central position (Fig. 1a,b). The three Gs at the 3′ portion of the SD sequence, in conjunction with the following two nucleotides, base pair with the anti–Shine-Dalgarno (ASD) sequence in the presence of SAM24, which hinders the binding of 30S ribosomal subunits to the mRNA35. Mutational studies demonstrated that the SMK box riboswitch differs from most metabolite binding riboswitches in that the SD-ASD pairing is required for SAM binding, indicating that the output domain (the SD-ASD pairing) is an intrinsic part of the ligand binding domain24. Similarly to other SAM riboswitches, the SMK box riboswitch shows at least 100-fold preference for SAM over SAH (refs. 24,35 and this study).

Figure 1.

Figure 1

Structure of the SMK box riboswitch. (a) Secondary structure of the SMK box riboswitch based on phylogenetic analysis24. Black capital letters, residues that are 100% conserved; blue, 50–85% conserved; R, G or A; W, A or U; Y, C or U; thick dotted lines, hypervariable loops; solid ladders, conserved secondary structure; gray and pink shaded areas, SD sequence and AUG start codon, respectively. (b) Secondary structure of the SMK6 riboswitch RNA based on the crystal structure. Helices P1 through P4 are colored in cyan, green, silver and yellow, respectively. Gray shading, SD sequence; solid magenta lines, direct contacts between the RNA and the SAM molecule; dashed magenta lines, tertiary interactions between J3/2 and P2 and J2/4. Numbering is consistent with previous studies24,35. (c) Cartoon representation of the crystal structure of the SMK riboswitch. SAM is shown in overlapping CPK and surface representations in magenta and silver, respectively. The coloring scheme for the RNA is consistent with b.

To elucidate the SAM-induced translation-inhibition mechanism in the SMK box riboswitch, we determined crystal structures of the E. faecalis SMK box riboswitch, including its SD sequence, in the presence of SAM, selenium-derivitized SAM (Se-SAM) or SAH. The RNA was found to organize into a Y-shaped molecule, with SAM intercalated into the three-way helical junction (Fig. 1c). In agreement with previous genetic and enzymatic probing analyses24,35, the SD sequence contributes to both binding-site formation and specific SAM recognition. The near-cognate ligand SAH can make only weak and nonspecific interactions with the riboswitch.

RESULTS

Structure determination

We determined the crystal structure of the E. faecalis SMK box riboswitch24 to elucidate the SAM-dependent translation-inhibition mechanism. The optimized version of the SMK box riboswitch (Fig. 1b and Supplementary Methods online), designated SMK6, allowed structure determination at 2.2 Å and showed SAM binding activity similar to that of the full-length wild-type construct (Fig. 2a,b). The apparent Kd values for SAM determined by size-exclusion filtration were estimated to be 0.85 µM for the full-length transcript (corresponding to positions 15–118 relative to the predicted E. faecalis metK transcription start site)24 and 0.57 µM for the 53-nucleotide (nt) crystallization construct (Fig. 2b). The Kd value for the full-length SMK transcript was further confirmed by fluorescence-quenching assays (Supplementary Fig. 1 and Supplementary Results online). In a competition assay, the SMK box riboswitch shows at least 100-fold preference for SAM over its near-cognate ligand SAH24 (Fig. 2b).

Figure 2.

Figure 2

Effects of SMK box riboswitch mutagenesis on SAM binding and apparent SAM binding constant (Kd) determination. (a) Effects of mutation (with s.d.) on SAM binding by size-exclusion filtration assay, as described24,35. (b) SAM and SAH Kd determination by competition binding assay. Binding curves are shown for the full-length E. faecalis metK RNA (open diamonds, Kd = 0.85 µM) and the SMK6 crystallization construct (filled diamonds, Kd = 0.57 µM). Unlabeled SAH showed no competition at a > 100-fold excess of SAH over SAM.

Overall structure of the SMK box riboswitch

The 53-nt SMK box riboswitch RNA folds into an inverted Y-shaped molecule, where helices P1 (SD-ASD helix) and P4 (linker helix) constitute the two short arms, and P3 (top helix) stacks on top of P2 (middle helix) to give rise to the long arm (Fig. 1c). Most of the secondary-structure features match the predictions from the phylogenetic analysis and RNase T1, V1, A and H probing experiments24, including the SAM-dependent formation of the SD-ASD helix and the protection of nucleotides including U69, U70, C75 and G88, although the P4 linker helix was not previously assigned. The quintuple-G SD sequence spans the SD-ASD (P1) and linker (P4) helices (Fig. 1b), contributing to the overall folding of the RNA and the SAM binding site. Among the 23 residues in the SMK box riboswitch that are 100% conserved (Fig. 1a), 10 participate in the formation of a pocket inside the three-way junction, where the adenosine moiety of SAM intercalates to allow continuous base stacking from P1 to P2. A lid-like structure formed as a result of a double-strand reversal in the J3/2 bulge partially encloses the SAM binding site and further stabilizes the three-way junction through base-triple interactions that enlarge the major groove of P2. A total of 15 divalent metal ions were identified bound to the SMK RNA to shield the unfavorable electrostatic interactions between sugar phosphate backbones as the result of tertiary RNA folding (Supplementary Fig. 2 and Supplementary Table 1 online).

The P4 linker helix was further probed by poly(A) scanning mutations. Replacement of the 11-nt hypervariable E. faecalis linker region with a poly(A)6 sequence did not affect SAM binding activity; disrupting the linker helix by further mutating G88 and G89 to adenosine residues, however, dramatically reduced SAM binding (Fig. 2a), consistent with the participation of these residues in the P4 linker helix. Additional point mutations were generated to confirm the formation of the G88-C76 base pair (Fig. 2a). Introduction of a C76U wobble mutation in the E. faecalis SMK box, which occurs naturally in some SMK box family members when a third 77–87 base pair is maintained24, resulted in retention of 59% of the SAM binding activity. By contrast, the G88A mismatch mutation that would severely destabilize the linker helix resulted in a 75-fold reduction in SAM binding activity. Combination of the C76U and G88A mutations, which replaces the G-C pair with a weaker A-U pair, restored the binding to 32% of the wild-type level, consistent with sequence-alignment results that suggest that an additional base pair is needed in this context to maintain the stability of the linker helix24.

Binding-pocket formation

The floor of the SAM binding pocket (Fig. 3a) is defined by a crucial base-triple interaction (A73•G90-C25), where N1 of A73 approaches from the minor groove side to accept a hydrogen bond from N2 of G90 (Fig. 3b). This base triple serves two purposes: first, it ties J2/4 to the P1 SD-ASD helix in conjunction with another interaction between the N1 of A74 and the 2′-hydroxyl of G90; furthermore, it orients the N6 amine of A73 for SAM recognition. Base-pairing at C25-G90 is crucial, as a C25•A90 mismatch (G90A mutation) reduced the SAM binding activity to 1.3% of the wild type, whereas a U25-G90 wobble pair (C25U mutation) showed 63% of the SAM binding activity (Fig. 2a). The importance of the A73•G90-C25 base triple is underlined by the strong, deleterious effect of the C25U G90A double mutation, which converts the SD element from the quintuple-G sequence found in all SMK box elements to the more commonly found GGAGG sequence, while maintaining the SD-ASD pairing. The loss of SAM binding activity in this mutant is apparently due to the loss of N2 when G90 is converted to A, which disrupts the A73•G90 side of the base-triple interaction. Thus, the structural and mutational data collectively explain the absolute conservation of an unusual SD sequence in all SMK box riboswitch RNAs.

Figure 3.

Figure 3

SAM binding pocket and important interactions. The labeling and base coloring scheme are consistent with that in Figure 1. (a) Stereo view of the SAM binding site in the SMK box riboswitch. The adenosine moiety of SAM is shown to base-stack between U72 and G90. (b) The A73•G90-C35 base triple paves the ‘floor’ of the SAM binding pocket. The C-G base pair is co-planar, whereas A73 contacts from the minor groove of G90 at a 45° tilted angle, which orients N6 of A73 for SAM recognition one base plane above. (c) The A27•G71•G66 base triple defines the ‘ceiling’ of the binding pocket, where the sheared A27•G71 base pair in P2 is contacted at the major groove side by G66 of the J3/2 bulge. (d) Recognition of the adenosine base of SAM overlapped with the 2.2-Å experimental electron-density contoured at 1.5 σ. Atoms beyond sulfur in SAM do not have corresponding electron density, indicating disorder. The adenine of SAM is extensively recognized through hydrogen-bond interactions from A73 and G26. SAM has two intramolecular electrostatic interactions from the sulfonium ion to the O4′ and N3 of SAM. Grey mesh, RNA density; orange mesh, SAM density. (e) The 3′-hydroxyl group of the ribose of SAM is recognized by a hydrogen bond to the phosphoryl oxygen of G89, whereas the 2′-OH makes a rare π–hydrogen bond interaction with the N7 of G89. (f) The positively charged sulfonium ion of SAM makes favorable electrostatic interactions with O4 of U72 and the 2′-hydroxyl group of G71. Distances are given in angstroms. Carbon, oxygen, nitrogen, sulfur and phosphorus atoms are colored gray, red, blue, yellow and orange, respectively.

The ceiling of the SAM binding site is formed from two layers of nonstandard base pairs, both of which involve bases from the J3/2 trinucleotide bulge, which are highly conserved among the E. faecalis subfamily of the SMK box RNAs24. A sheared U72•A64 pair sits immediately on top of the SAM intercalation base plane, which exposes the O4 of U72 for SAM recognition. Further above is the A27•G71•G66 base triple, where G71 mediates extensive hydrogen bond contacts to the Hoogsteen face of both A27 and G66, weaving P2 together with J3/2 (Fig. 3c). U65 in the J3/2 bulge mediates a U-turn motif and extends toward the opening of the SAM binding pocket; this residue was shown to be more prone to RNase T1 digestion in the presence of SAM24, consistent with the proposal that SAM binding reorients this residue. The backbone of the bulge resembles a ‘lid’ that secludes SAM inside the three-way junction and contributes to the overall electronegative environment inside the pocket, which may help to attract the SAM molecule. The importance of the J3/2 lid structure is indicated by the U65C and U65 deletion mutations, which cause a 5-fold and 20-fold decrease in SAM binding activity, respectively (Fig. 2a).

SAM recognition

The universally conserved G26 is left unpaired inside the cavity of the three-way junction. SAM intercalates its adenosine moiety between the P1 and P2 helices from the major groove side, stabilizing the three-way junction through π-stacking interactions (Fig. 3d). The adenosine moiety of SAM adopts an energetically unfavorable syn-conformation and presents its Watson-Crick face to form three minor groove contacts (N1-amino, amino-N3 and amino-2′-OH) with the unpaired G26. Even a minor perturbation of the G26-SAM interaction, such as occurs in the G26A mutant, in which the N1-amino hydrogen bond is disrupted, reduces SAM binding by five-fold (Fig. 2a). The N7 at the Hoogsteen face accepts a hydrogen bond from the N6 amine of the tilted A73 (Fig. 3d). Disruption of this hydrogen bond by the A73G mutation caused an 80-fold reduction in SAM binding activity (Fig. 2a), although we cannot completely separate this effect from the disruption of the A73•G90-C25 base triple. The ribose moiety of SAM adopts a 2′-endo conformation to avoid steric clashes with the base plane of G89 in the SD sequence36. The 2′- and 3′-hydroxyl groups of SAM are recognized by hydrogen bonds to the N7 (π-hydrogen bond) and 2′-hydroxyl of G89, respectively (Fig. 3e). The 2′-hydroxyl of SAM donates an additional water-mediated hydrogen bond to the N6 of A74.

The positive charge on the sulfonium ion of SAM is recognized through favorable electrostatic interactions with the O4 carbonyl of U72 and the 2′-hydroxyl of G71 (Fig. 3f). The positive charge is further stabilized through intramolecular electrostatic interactions with the O4′ on the ribose and N3 on the base of SAM (Fig. 3d). Consistent with the structural observations, the U72C mutation, which places a partially positive N4 amine toward the sulfonium, decreases the SAM binding activity by 70-fold (Fig. 2a). A similar charge-stabilization scheme involving contacts between the sulfonium ion of SAM and O4 of uracil is used by the other two classes of SAM binding riboswitches to selectively bind the SAM molecule33,34. The neutral, hydrophobic sulfide in SAH is not expected to make these electrostatic interactions with the RNA.

In contrast to the well-defined adenosine moiety, no electron density was observed for functional groups beyond the sulfonium ion in SAM (Fig. 3d). This is a strong indication that the methionine tail (including the main chain and most of the side chain atoms beyond the sulfur) is not specifically recognized by the SMK box riboswitch, which is in sharp contrast to the extensive recognition of the methionine tail observed in the structures of the S box and SAM-II riboswitches33,34. Lack of recognition toward the methionine tail is further supported by binding studies with SAM analogs, where SAM analogs lacking the methionine main chain atoms bind the SMK box riboswitch with affinities similar to that of SAM (A.M.S., F.J. Grundy and T.M.H., unpublished data). The methyl group on the sulfonium ion is not specified by the SMK box riboswitch (Fig. 3d) because replacing it with an ethyl group does not impair ligand-RNA interactions (A.M.S., F.J. Grundy and T.M.H., unpublished data). The modeled methionine conformation illustrated in Figure 3 complies with all spatial constraints derived from the structural and SAM-analog studies.

Regulation of E. faecalis metK-lacZ fusions in vivo

The E. faecalis metK SMK box was previously shown to confer translational regulation of a lacZ reporter gene in which the translation-initiation region is replaced by that of the metK gene24. Integration of the fusion construct into the B. subtilis chromosome results in high expression when cells are grown under conditions where intracellular SAM pools are low and a five-fold reduction in expression when cells are grown in the presence of high methionine concentrations, conditions that result in high SAM pools37.

The SMK box riboswitch structure supports the model that SAM-dependent SD-ASD helix formation is responsible for translational inhibition. To verify this hypothesis, we evaluated the functional consequences of three sets of mutations that perturb the stability of the SD-ASD helix. As previously reported24, a C24G mutation in the ASD region, predicted to disrupt the SD-ASD interaction, causes loss of repression in vivo (Table 1) and loss of SAM binding (Fig. 2a). A C24U mutation had a similar effect (Table 1), although expression was significantly higher and a small response to SAM was retained, presumably because of the maintenance of a U24-G91 wobble pairing. A C24U G91A double mutation, which restores pairing at this position, restored SAM binding (Fig. 2a) and resulted in low expression but only partial restoration of repression by SAM (Table 1). The low expression is likely to be due to disruption of the SD by the G91A mutation, which causes reduced affinity for 30S ribosomal subunits and therefore obscures the regulatory response in vivo. A similar pattern was observed with the U22G and A93C mutations (Table 1), which affect the ASD and SD sequences, respectively, and cause loss of SAM binding in vitro35. The U22G A93C double mutant restores the pairing at this position and restores SAM binding35. This mutant also showed SAM-dependent repression in vivo, although expression was reduced relative to that of the wild-type construct (Table 1). This reduction in expression may be due to an enhanced SD-ASD interaction from replacing an A-U pair with a G-C pair.

Table 1.

Expression of E. faecalis metK-lacZ fusions

−Met +Met Ratiob
Wild type 120 ± 19a   24 ± 5.5 5.0
C24G 110 ± 14   95 ± 6.1 1.2
C24U 310 ± 0.71 200 ± 22 1.6
C24U+G91A   23 ± 2.9   15 ± 2.9 1.5
U22G 220 ± 27 160 ± 13 1.4
A93C   77 ± 8.7   42 ± 1.0 1.8
U22G+A93C   11 ± 3.7  2.5 ± 0.72 4.4
C25U 225 ± 25   80 ± 5.0 2.8
C25U+G90A   78 ± 19   82 ± 16 0.95
a

β-Galactosidase activities are expressed in Miller units54.

b

Ratio of expression during growth in the absence of methionine to expression during growth in the presence of methionine.

As described above, G90 in the SD sequence has a pivotal role in the SMK box riboswitch, as it organizes the SAM binding pocket through A73•G90-C25 base-triple formation and stacks directly underneath the SAM molecule. The C25U substitution creates a wobble pair with G90. This mutant showed partial repression by high SAM concentrations in vivo (Table 1), consistent with a modest reduction of SAM binding in vitro (Fig. 2a). The C25U G90A double mutant is predicted to restore Watson-Crick pairing at this position but not the crucial base-triple interaction with A73. This variant showed complete loss of repression in vivo, despite maintaining a canonical SD sequence (Table 1), in accordance with the loss of SAM binding in vitro (Fig. 2a). These data confirm the importance of G90 for SAM binding and SAM-dependent regulation.

Se-SAM–bound SMK box structure confirms sulfonium position

To confirm the position of the sulfur atom in the SAM molecule, we further determined the structure of the SMK box riboswitch soaked in freshly prepared Se-SAM (where the sulfur atom is substituted with selenium, a heavier chalcogen analog38). The conformations of the RNA and the Se-SAM molecule are essentially identical to that in the SAM-bound structure, with an r.m.s. deviation of 0.3 Å for all-phosphorous-atom alignment. The same set of contacts specifies Se-SAM in the ligand binding site, and the selenomethionine moiety remains unstructured beyond the onium selenium (Fig. 4a). The location of the selenium atom, which essentially overlaps with the sulfur atom in the omit map, is unambiguously identified by an 8σ anomalous difference signal collected at the absorption edge of selenium (Fig. 4a). The distances of the electrostatic interactions between the RNA and the onium selenium are on average ~0.3 Å longer than seen in the SAM-bound structure, reflecting the increased van der Waals radius. Observations from the Se-SAM–bound riboswitch structure confirm our assignment of the sulfonium position in SAM and strengthen our conclusion that the methionine tail of SAM is not specified by the SMK box riboswitch.

Figure 4.

Figure 4

Binding of Se-SAM and SAH to the SMK box riboswitch. (a) Stereo view of the binding pocket in the SMK box riboswitch in complex with Se-SAM. The location of the selenium atom is confirmed by the strong anomalous-difference density shown in blue contoured at 8 σ. The rest of the binding pocket in the Se-SAM–bound SMK structure is almost identical to that in the SAM-bound structure. Magenta mesh signifies the simulated composite omit electron-density map of Se-SAM contoured at 1.5 σ. (b) Stereo view of the SAH-bound SMK structure from a direction similar to that shown in a. The simulated annealing omit map contoured at 0.8 σ level clearly shows that the ribose and sulfide moieties rotate 180° to exit the RNA from the linker helix side.

SAH makes a minimum set of contacts to the SMK box riboswitch

To investigate how the SMK box riboswitch would respond in the presence of natural near-cognate ligands such as SAH, we determined the SMK box RNA structure in the presence of SAH at a saturating concentration of 2 mM, about 67- to 2,000-fold higher than the intracellular concentration of SAH (0.5–30 µM39). The resulting structure revealed that SAH can displace SAM at saturating concentration. The binding pocket and overall structure of the SMK box riboswitch remains unchanged, presumably owing to strong crystal lattice contacts to the P1 and P4 helices. The resulting structure is thus a snapshot of the functional SMK box riboswitch sampling a near-cognate ligand before rejecting it. In the 2.9-Å resolution structure, SAH adopts a conformation very different from SAM in the binding pocket (Fig. 4b). The ribose moiety of SAH rotates 180°, such that the adenosine moiety now adopts an anti-conformation. 2′- and 3′-hydroxyls in SAH make an alternative set of interactions with the phosphoryl oxygens of G90. More importantly, weak electron-density features revealed that the uncharged sulfide in SAH rotates 180° from U72 to the vicinity of G71. As a result, the intramolecular contacts that stabilize the sulfonium ion conformation in SAM and the key electrostatic interactions between the sulfur atom and the RNA are all lost when SAH is in place. It therefore seems that the positive charge on the sulfur atom is crucial in presenting the SAM molecule in the correct conformation for strong ligand-RNA interactions, analogous to the ‘lock-and-key’ mechanism found in many enzyme-substrate complexes. The structural observations are consistent with in vitro assays showing that the SMK box riboswitch has much higher affinity for SAM over SAH24,35, with at least a 100-fold preference for SAM.

DISCUSSION

With the completion of the structure of the SMK box riboswitch, a family portrait of the three known classes of SAM binding riboswitches is now available33,34. These riboswitches adopt completely different RNA folds and have probably emerged independently during evolution. They nevertheless converged at the functional level to preferentially recognize SAM, an important metabolite inside the cell, and regulate gene expression in response to the ligand binding event. We conclude from structural comparisons that, although creative ways are used to accommodate the binding of SAM, a conserved mechanism is used by all three classes of SAM riboswitches to distinguish SAM from SAH, which is the biologically relevant SAM analog.

Riboswitches sense both chemical and conformational distinctions in SAM

SAM adopts drastically different conformations among the three classes of SAM riboswitches (Supplementary Fig. 3a online). The Thermoanaerobacter tengcongensis and B. subtilis yitJ S box (SAM-I) riboswitches completely engulf SAM in a pocket between two helical stacks (ref. 33 and C.L. et al., unpublished data). The SAM-II riboswitch adopts a classic H-type pseudoknot and encloses SAM inside an RNA triple helix34. The SMK box riboswitch allows the intercalation of SAM into a tight pocket in a three-way junction. In each case, SAM mediates important tertiary interactions to stabilize the ligand-bound conformation of the riboswitch. The adenine base is involved in base stacking and base-triple interactions in all three riboswitches. The positively charged sulfonium moiety is invariably recognized through favorable electrostatic interactions, usually with one or two O4 carbonyl oxygen atoms from uracil residues. This recognition forms the basis for preferential binding of SAM over SAH. The methyl group on the sulfonium moiety is not directly contacted by any of the three riboswitches. Rather, it points toward the solvent region and is recognized collectively with the sulfonium ion, a mechanism perhaps evolved to prevent self inactivation by spontaneous methylation of the RNA by SAM. Recognition of the methionine tail, however, occurs differently in the three riboswitches. The tail is extensively contacted by the S box riboswitch33, less so by the SAM-II riboswitch34 and completely ignored by the SMK box riboswitch. Previous structural studies seem to suggest that riboswitches are more likely to be identified through bioinformatics rather than SELEX approaches because in each structure, the metabolite is completely encapsulated by the RNA. The ability of the biologically active SMK box riboswitch to recognize SAM by specifying only half of the molecule and without engulfing SAM completely suggests that additional classes of SAM riboswitches can potentially be identified by SELEX approaches using a SAM molecule immobilized at the methionine tail as the bait.

In addition to chemical differences between SAM and SAH, SAM also adopts a distinct conformation that is energetically unfavorable for SAH and other noncognate SAM analogs. Alignment of the conformations of SAM within the three riboswitch structures along the ribose region clearly revealed that, although the conformation of the adenine base (syn- or anti-) and the methionine tail (crouched or extended) varies dramatically among the three riboswitch structures, the sulfonium ion makes an invariable, strong 3-Å electrostatic contact to the O4′ of the SAM ribose (Supplementary Fig. 3a,b). This gauche-conformation (Supplementary Fig. 3b) about the C4′–C5′ bond has been shown by NMR measurement to be the predominant (93%) SAM conformation in solution36. The same study showed that SAH, on the other hand, favors the anti-conformation36, presumably because of its inability to maintain the intramolecular electrostatic interaction. Indeed, in our survey of 30 randomly chosen SAH-bound protein structures, 87% of the SAH molecules adopted the anti-conformation (Supplementary Fig. 3b). Thus, by simultaneously contacting both the ribose and the sulfonium moieties of SAM, all three riboswitches effectively select for a SAM conformation that is unfavorable for SAH. Consistent with this observation, our structural analysis of the SAH-bound SMK box riboswitch revealed that, although SAH could bind to the crystalline-trapped SMK box riboswitch at a nonphysiological high concentration, the sulfide moiety in the riboswitch-bound form of SAH swings 180° away from the sulfonium binding site, severely weakening the ligand-RNA interactions to a level indistinguishable from the binding of other adenosine-containing metabolites such as ATP. Thus, it seems that both chemical and conformational differences at the sugar-sulfur linkage are explored by SAM binding riboswitches to distinguish between cognate and near-cognate ligands. The recent discovery of a new class of SAH binding riboswitches upstream of bacterial genes involved in SAM recycling40 demonstrates that RNA is capable of forming a selective binding site that favors either SAM or SAH.

A simple SAM-dependent translation-inhibition mechanism

A typical riboswitch is conceptually modular, consisting of a ligand-sensing aptamer domain and a separate output domain whose structure is influenced by the ligand binding event at the aptamer domain. A much simpler mechanism for translational regulation is found in the SMK box and SAM-II riboswitches, where the SD sequence is an integral part of the SAM binding aptamer domain, allowing ligand sensing and translational inhibition through SD-sequence sequestration to take place in a single step. An important feature in the SMK box riboswitch is the direct involvement of the SD sequence in SAM recognition. Although strong crystal packing interactions prevent observation of large ligand-induced conformational changes in crystal structures, mutagenesis and enzymatic probing assays24,35 clearly revealed global conformational changes in the presence of physiologically relevant concentrations of SAM but not SAH. These SAM-dependent changes are especially evident in the formation of the linker and SD-ASD helices that sequester the ribosome binding site. Conversely, the ability to form the linker and SD-ASD helices is a prerequisite for SAM binding24. The combined structural and biochemical data suggest that SAM shifts the conformational equilibrium toward the ligand-bound state seen in the crystal structure, where the SD sequence required for binding of the 30S ribosomal subunit is sequestered. Consistent with this mechanism, ribosome toeprinting analysis showed that the correct positioning of the 30S ribosomal subunit on the SMK box mRNA is reduced in the presence of SAM but not SAH35. Direct involvement of the SD sequence in the formation of the aptamer domain is also observed in the SAM-II riboswitch structure34, although the exact location of the SD sequence is less well defined.

In summary, our combined structural and mutagenesis data on the SMK box riboswitch clearly demonstrate its mechanism for translational inhibition through sequestration of the SD sequence, and the interplay between SD-ASD pairing and SAM binding activity. The SMK box riboswitch is unique among riboswitch RNAs studied to date in that the same residues of the RNA that are involved in gene regulation (the SD sequence) are directly involved in specific recognition of the SAM ligand.

METHODS

RNA crystallization and structure determination

Design of the crystallization construct is described in the Supplementary Methods. RNA was prepared as described41. Se-SAM was a gift from S. Booker (Pennsylvania State University). Its synthesis has been described previously38. SAM and SAH were purchased from Sigma, dissolved in water and DMSO, respectively, as 100 mM stock solutions (pH < 7) and stored at −80 °C until immediately before use.

To ensure conformational homogeneity, the SMK box RNA was heat-refolded as described24,35. The high-resolution SMK6 crystals grew as extremely slim needle crystals of 15 × 15 µm2 in cross-section from a solution containing 40 mM sodium cacodylate, pH 7.0, 80 mM strontium chloride, 15% (w/v) 2-methylpentane-2,4-diol (MPD) and 2 mM spermine-HCl. The MPD content was raised stepwise to 25% (w/v) before the crystals were flash-frozen in liquid nitrogen. Diffraction data were collected using the microcrystallography setup at beamlines Advanced Photon Source (APS) 24-ID-E and MACCHESS F1 and processed using HKL200042. The detailed phasing and refinement procedure is described in the Supplementary Methods. The initial model was built using COOT43 from experimental phases calculated from SHELXD44 and refined using Refmac545 and CNS46,47. The final model includes all nucleotides and a total of 15 strontium ions (Supplementary Table 1) and 69 water molecules. The methionine tail of SAM and the pyrimidine ring of U65 are the only portion of the model without corresponding electron densities. Final Rwork and Rfree factors were 22.1% and 22.7%, respectively (Table 2). Se-SAM and SAH bound structures were determined by molecular replacement using PHASER48 and refined using Refmac5 (ref. 45) and CNS46,47 (Table 2 and Supplementary Methods).

Table 2.

Crystallographic statistics

SMK native bound
to SAM #1
SMK native bound
to SAM #2
SMK Ir derivative SMK bound to Se-SAM SMK bound to SAH
Data collection
Space group I 4122 P 41212 P 41212 I 4122 I 4122
Cell dimensions
    a, b, c (Å) 97.806, 97.806, 87.138 98.131, 98.131, 86.262 98.180, 98.180, 86.790 96.776, 96.776, 86.765 98.148, 98.148, 86.393
Resolution (Å) 20 (2.2) 20 (2.6) 20 (3.3) 20 (2.7) 20 (2.9)
Rsym 7.1 (30.3) 11.3 (40.1) 13.3 (42.2) 7.6 (28.4) 7.9 (24.6)
I / σI 29.4 (1.2) 12.3 (2.59) 13.2 (5.8) 15.5 (1.7) 80.8 (11.4)
Completeness (%) 95.1 (75.4) 90.4 (85.0) 100 (100) 93.7 (74.0) 100 (100)
Redundancy 9.4 (2.3) 6.3 (4.8) 13.7 (13.6) 3.9 (2.8) 9.3 (8.0)
Refinement
Resolution (Å) 2.2 2.7 2.9
No. reflections 9,147 4,980 4,694
Rwork / Rfree 22.1 / 22.7 22.7 / 25.5 22.2 / 25.9
No. atoms
    RNA 1,155 1,155 1,155
    Ligand/ion 42 43 31
    Water 69 11 5
B-factors 53 57 56
    RNA 50 57 57
    Ligand/ion 55 62 49
    Water 53 45 44
R.m.s. deviations
    Bond lengths (Å) 0.008 0.005 0.006
    Bond angles (°) 1.61 1.29 1.43
*

Values in parentheses are for highest-resolution shell.

In vitro SAM binding assays

We carried out SAM binding assays as previously described24,35 with minor modifications. DNA templates corresponding to positions 15–118 (relative to the predicted transcription start site) of the E. faecalis metK leader were constructed by ligating overlapping pairs of complementary oligonucleotides, including a T7 RNA polymerase promoter sequence, as previously described49,50. Ligated products were amplified by PCR and RNAs were synthesized by T7 RNA polymerase transcription using an AmpliScribe T7 High Yield Transcription Kit (Epicentre Biotechnologies). RNAs (3 µM) in 1× transcription buffer51 were heated to 65 °C for 5 min and slow-cooled to 40 °C, followed by addition of radiolabeled SAM (3 µM [methyl-3H]-SAM, 15 Ci mmol−1; GE Healthcare) in a total reaction volume of 40 µl. Binding reactions were incubated at room temperature (20–25 °C) for 15 min followed by passage through a Nanosep 10K Omega filter (Pall Life Sciences) by centrifugation at 14,000g for 2.5 min. Filters were washed four times with 40 µl 1× transcription buffer to remove unbound SAM, and material retained by the filter was collected, mixed with Packard BioScience Ultima Gold scintillation fluid and counted in a Packard Tri-Carb 2100TR liquid scintillation counter. No SAM was retained by the filter in the absence of RNA; hence, the background for nonspecific binding is negligible. All binding assays were carried out in triplicate and the s.d. are shown in Figure 2.

Apparent Kd determination

The apparent equilibrium dissociation constants (Kd values) for the wild-type 15–118 metK leader sequence and the 53-nucleotide crystallization construct were determined using a modified SAM binding assay37. Briefly, T7 RNA polymerase transcribed RNA (1 µM) in 1× transcription buffer51 was refolded as described above and incubated with [3H]-SAM ranging in concentration from 0.02 µM to 10 µM. Binding reactions were loaded onto Nanosep 3K Omega filters (Pall Life Sciences) and centrifuged briefly at 14,000g to avoid large volume changes. The concentration of the RNA-[3H]-SAM complex retained by the filter and the unbound SAM present in the flow through were determined by scintillation counting of known volumes, and nonlinear regression analyses were performed using Kaleida-Graph Version 3.51 (Synergy Software). Apparent Kd values represent the averages of at least two independent experiments for each construct with a margin of error ≤ 5%.

In vivo lacZ reporter gene assay of SMK box fusions

We carried out in vivo lacZ reporter gene assays as described previously24,35. Enterococcus faecalis metK leader region constructs, which included the first 15 nt of the coding region, were positioned downstream of the highly expressed B. subtilis glyQS promoter. The resulting DNA fragment was inserted into a lacZ fusion vector (pFG328)52 to generate an in-frame metK-lacZ translational fusion in which the first five codons of metK were fused to codon 18 of lacZ. The constructs were introduced into the chromosome of B. subtilis strain BR151 (metB10 lys-3 trpC2) by integration into an SPβ prophage. Bacillus subtilis strains containing the lacZ fusions were grown at 37 °C in Spizizen minimal medium53 containing methionine (50 µg ml−1) until early exponential growth. The cells were harvested by centrifugation at 8,000g and resuspended in Spizizen medium either with or without methionine, and samples were collected at 1-h intervals and assayed for β-galactosidase activity after toluene permeabilization of the cells54. All assays were carried out in triplicate and the s.d. are shown in Table 1.

Accession codes

Protein Data Bank: Structure factors and coordinates for the SAM, SAH and Se-SAM bound SMK box riboswitch have been deposited in the Protein Data Bank with accession numbers 3E5C, 3E5E and 3E5F, respectively.

Supplementary Material

supplmental_data_all

ACKNOWLEDGMENTS

We thank S. Booker at Pennsylvania State University for the gift of the Se-SAM; Q. Hao, R. Gillian and other beam line staff at MACCHESS and APS ID24 for assistance in data collection; F. Grundy for assistance in SAM binding assay design; and J. Doudna for helpful discussions and comments on the manuscript. Work in the Henkin laboratory was supported by a Public Health Service grant (GM63615) from the US National Institutes of Health (NIH). This work is based upon research conducted at the Northeastern Collaborative Access Team beamlines of the Advanced Photon Source, which is supported by National Cancer Institute (NCI) award RR-15301 and DOE contract No. DE-AC02-06CH11357, and at the Cornell High Energy Synchrotron Source (CHESS), which is supported by the National Science Foundation award DMR 0225180 and the NCI award RR-01646.

Footnotes

Note: Supplementary information is available on the Nature Structural & Molecular Biology website.

References

  • 1.Tucker BJ, Breaker RR. Riboswitches as versatile gene control elements. Curr. Opin. Struct. Biol. 2005;15:342–348. doi: 10.1016/j.sbi.2005.05.003. [DOI] [PubMed] [Google Scholar]
  • 2.Grundy FJ, Henkin TM. From ribosome to riboswitch: control of gene expression in bacteria by RNA structural rearrangements. Crit. Rev. Biochem. Mol. Biol. 2006;41:329–338. doi: 10.1080/10409230600914294. [DOI] [PubMed] [Google Scholar]
  • 3.Vitreschak AG, Rodionov DA, Mironov AA, Gelfand MS. Riboswitches: the oldest mechanism for the regulation of gene expression? Trends Genet. 2004;20:44–50. doi: 10.1016/j.tig.2003.11.008. [DOI] [PubMed] [Google Scholar]
  • 4.Winkler W, Nahvi A, Breaker RR. Thiamine derivatives bind messenger RNAs directly to regulate bacterial gene expression. Nature. 2002;419:952–956. doi: 10.1038/nature01145. [DOI] [PubMed] [Google Scholar]
  • 5.Wickiser JK, Winkler WC, Breaker RR, Crothers DM. The speed of RNA transcription and metabolite binding kinetics operate an FMN riboswitch. Mol. Cell. 2005;18:49–60. doi: 10.1016/j.molcel.2005.02.032. [DOI] [PubMed] [Google Scholar]
  • 6.Serganov A, Polonskaia A, Phan AT, Breaker RR, Patel DJ. Structural basis for gene regulation by a thiamine pyrophosphate-sensing riboswitch. Nature. 2006;441:1167–1171. doi: 10.1038/nature04740. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Winkler WC, Cohen-Chalamish S, Breaker RR. An mRNA structure that controls gene expression by binding FMN. Proc. Natl. Acad. Sci. USA. 2002;99:15908–15913. doi: 10.1073/pnas.212628899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Sudarsan N, Cohen-Chalamish S, Nakamura S, Emilsson GM, Breaker RR. Thiamine pyrophosphate riboswitches are targets for the antimicrobial compound pyrithiamine. Chem. Biol. 2005;12:1325–1335. doi: 10.1016/j.chembiol.2005.10.007. [DOI] [PubMed] [Google Scholar]
  • 9.Thore S, Leibundgut M, Ban N. Structure of the eukaryotic thiamine pyrophosphate riboswitch with its regulatory ligand. Science. 2006;312:1208–1211. doi: 10.1126/science.1128451. [DOI] [PubMed] [Google Scholar]
  • 10.Christiansen LC, Schou S, Nygaard P, Saxild HH. Xanthine metabolism in Bacillus subtilis: characterization of the xpt-pbuX operon and evidence for purine- and nitrogen-controlled expression of genes involved in xanthine salvage and catabolism. J. Bacteriol. 1997;179:2540–2550. doi: 10.1128/jb.179.8.2540-2550.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Mandal M, Breaker RR. Adenine riboswitches and gene activation by disruption of a transcription terminator. Nat. Struct. Mol. Biol. 2004;11:29–35. doi: 10.1038/nsmb710. [DOI] [PubMed] [Google Scholar]
  • 12.Mandal M, Boese B, Barrick JE, Winkler WC, Breaker RR. Riboswitches control fundamental biochemical pathways in Bacillus subtilis and other bacteria. Cell. 2003;113:577–586. doi: 10.1016/s0092-8674(03)00391-x. [DOI] [PubMed] [Google Scholar]
  • 13.Roth A, et al. A riboswitch selective for the queuosine precursor preQ1 contains an unusually small aptamer domain. Nat. Struct. Mol. Biol. 2007;14:308–317. doi: 10.1038/nsmb1224. [DOI] [PubMed] [Google Scholar]
  • 14.Kim JN, Roth A, Breaker RR. Guanine riboswitch variants from Mesoplasma florum selectively recognize 2′-deoxyguanosine. Proc. Natl. Acad. Sci. USA. 2007;104:16092–16097. doi: 10.1073/pnas.0705884104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Meyer MM, Roth A, Chervin SM, Garcia GA, Breaker RR. Confirmation of a second natural preQ1 aptamer class in Streptococcaceae bacteria. RNA. 2008;14:685–695. doi: 10.1261/rna.937308. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Mandal M, et al. A glycine-dependent riboswitch that uses cooperative binding to control gene expression. Science. 2004;306:275–279. doi: 10.1126/science.1100829. [DOI] [PubMed] [Google Scholar]
  • 17.Rodionov DA, Vitreschak AG, Mironov AA, Gelfand MS. Regulation of lysine biosynthesis and transport genes in bacteria: yet another RNA riboswitch? Nucleic Acids Res. 2003;31:6748–6757. doi: 10.1093/nar/gkg900. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Sudarsan N, Wickiser JK, Nakamura S, Ebert MS, Breaker RR. An mRNA structure in bacteria that controls gene expression by binding lysine. Genes Dev. 2003;17:2688–2697. doi: 10.1101/gad.1140003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Grundy FJ, Lehman SC, Henkin TM. The L box regulon: lysine sensing by leader RNAs of bacterial lysine biosynthesis genes. Proc. Natl. Acad. Sci. USA. 2003;100:12057–12062. doi: 10.1073/pnas.2133705100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Winkler WC, Nahvi A, Sudarsan N, Barrick JE, Breaker RR. An mRNA structure that controls gene expression by binding S-adenosylmethionine. Nat. Struct. Biol. 2003;10:701–707. doi: 10.1038/nsb967. [DOI] [PubMed] [Google Scholar]
  • 21.McDaniel BA, Grundy FJ, Artsimovitch I, Henkin TM. Transcription termination control of the S box system: direct measurement of S-adenosylmethionine by the leader RNA. Proc. Natl. Acad. Sci. USA. 2003;100:3083–3088. doi: 10.1073/pnas.0630422100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Epshtein V, Mironov AS, Nudler E. The riboswitch-mediated control of sulfur metabolism in bacteria. Proc. Natl. Acad. Sci. USA. 2003;100:5052–5056. doi: 10.1073/pnas.0531307100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Corbino KA, et al. Evidence for a second class of S-adenosylmethionine riboswitches and other regulatory RNA motifs in α-proteobacteria. Genome Biol. 2005;6:R70. doi: 10.1186/gb-2005-6-8-r70. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Fuchs RT, Grundy FJ, Henkin TM. The SMK box is a new SAM-binding RNA for translational regulation of SAM synthetase. Nat. Struct. Mol. Biol. 2006;13:226–233. doi: 10.1038/nsmb1059. [DOI] [PubMed] [Google Scholar]
  • 25.Groisman EA, Cromie MJ, Shi Y, Latifi TA. Mg2+-responding RNA that controls the expression of a Mg2+ transporter. Cold Spring Harb. Symp. Quant. Biol. 2006;71:251–258. doi: 10.1101/sqb.2006.71.005. [DOI] [PubMed] [Google Scholar]
  • 26.Coppins RL, Hall KB, Groisman EA. The intricate world of riboswitches. Curr. Opin. Microbiol. 2007;10:176–181. doi: 10.1016/j.mib.2007.03.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Winkler WC. Metabolic monitoring by bacterial mRNAs. Arch. Microbiol. 2005;183:151–159. doi: 10.1007/s00203-005-0758-9. [DOI] [PubMed] [Google Scholar]
  • 28.Sudarsan N, Barrick JE, Breaker RR. Metabolite-binding RNA domains are present in the genes of eukaryotes. RNA. 2003;9:644–647. doi: 10.1261/rna.5090103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Cheah MT, Wachter A, Sudarsan N, Breaker RR. Control of alternative RNA splicing and gene expression by eukaryotic riboswitches. Nature. 2007;447:497–500. doi: 10.1038/nature05769. [DOI] [PubMed] [Google Scholar]
  • 30.Henkin TM, Yanofsky C. Regulation by transcription attenuation in bacteria: how RNA provides instructions for transcription termination/antitermination decisions. Bioessays. 2002;24:700–707. doi: 10.1002/bies.10125. [DOI] [PubMed] [Google Scholar]
  • 31.Grundy FJ, Henkin TM. The S box regulon: a new global transcription termination control system for methionine and cysteine biosynthesis genes in Gram-positive bacteria. Mol. Microbiol. 1998;30:737–749. doi: 10.1046/j.1365-2958.1998.01105.x. [DOI] [PubMed] [Google Scholar]
  • 32.Barrick JE, Breaker RR. The distributions, mechanisms, and structures of metabolite-binding riboswitches. Genome Biol. 2007;8:R239. doi: 10.1186/gb-2007-8-11-r239. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Montange RK, Batey RT. Structure of the S-adenosylmethionine riboswitch regulatory mRNA element. Nature. 2006;441:1172–1175. doi: 10.1038/nature04819. [DOI] [PubMed] [Google Scholar]
  • 34.Gilbert SD, Rambo RP, Van Tyne D, Batey RT. Structure of the SAM-II riboswitch bound to S-adenosylmethionine. Nat. Struct. Mol. Biol. 2008;15:177–182. doi: 10.1038/nsmb.1371. [DOI] [PubMed] [Google Scholar]
  • 35.Fuchs RT, Grundy FJ, Henkin TM. S-adenosylmethionine directly inhibits binding of 30S ribosomal subunits to the SMK box translational riboswitch RNA. Proc. Natl. Acad. Sci. USA. 2007;104:4876–4880. doi: 10.1073/pnas.0609956104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Stolowitz ML, Minch MJ. S-adenosyl-L-methionine and 5′-adenosyl-L-homocysteine, an NMR Study. J. Am. Chem. Soc. 1981;103:6015–6019. [Google Scholar]
  • 37.Tomsic J, McDaniel BA, Grundy FJ, Henkin TM. Natural variability in SAM-dependent riboswitches: S box elements in Bacillus subtilis exhibit differential sensitivity to SAM in vivo and in vitro. J. Bacteriol. 2007;190:823–833. doi: 10.1128/JB.01034-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Iwig DF, Booker SJ. Insight into the polar reactivity of the onium chalcogen analogues of S-adenosyl-L-methionine. Biochemistry. 2004;43:13496–13509. doi: 10.1021/bi048693+. [DOI] [PubMed] [Google Scholar]
  • 39.Ueland PM. Pharmacological and biochemical aspects of S-adenosylhomocysteine and S-adenosylhomocysteine hydrolase. Pharmacol. Rev. 1982;34:223–253. [PubMed] [Google Scholar]
  • 40.Wang JX, Lee ER, Morales DR, Lim J, Breaker RR. Riboswitches that sense S-adenosylhomocysteine and activate genes involved in coenzyme recycling. Mol. Cell. 2008;29:691–702. doi: 10.1016/j.molcel.2008.01.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Ke A, Doudna JA. Crystallization of RNA and RNA-protein complexes. Methods. 2004;34:408–414. doi: 10.1016/j.ymeth.2004.03.027. [DOI] [PubMed] [Google Scholar]
  • 42.Otwinowski Z, Minor W. Processing of X-ray diffraction data collected in oscillation mode. Methods Enzymol. 1997;276:307–326. doi: 10.1016/S0076-6879(97)76066-X. [DOI] [PubMed] [Google Scholar]
  • 43.Emsley P, Cowtan K. Coot: model-building tools for molecular graphics. Acta Crystallogr. D Biol. Crystallogr. 2004;60:2126–2132. doi: 10.1107/S0907444904019158. [DOI] [PubMed] [Google Scholar]
  • 44.Sheldrick GM. A short history of SHELX. Acta Crystallogr. A. 2008;64:112–122. doi: 10.1107/S0108767307043930. [DOI] [PubMed] [Google Scholar]
  • 45.Collaborative Computational Project, Number 4. The CCP4 suite: programs for protein crystallography. Acta Crystallogr. D Biol. Crystallogr. 1994;50:760–763. doi: 10.1107/S0907444994003112. [DOI] [PubMed] [Google Scholar]
  • 46.Brunger AT, et al. Crystallography NMR system: a new software suite for macromolecular structure determination. Acta Crystallogr. D Biol. Crystallogr. 1998;54:905–921. doi: 10.1107/s0907444998003254. [DOI] [PubMed] [Google Scholar]
  • 47.Brunger AT. Version 1.2 of the crystallography and NMR system. Nat. Protoc. 2007;2:2728–2733. doi: 10.1038/nprot.2007.406. [DOI] [PubMed] [Google Scholar]
  • 48.Storoni LC, McCoy AJ, Read RJ. Likelihood-enhanced fast rotation functions. Acta Crystallogr. D Biol. Crystallogr. 2004;60:432–438. doi: 10.1107/S0907444903028956. [DOI] [PubMed] [Google Scholar]
  • 49.Yousef MR, Grundy FJ, Henkin TM. tRNA requirements for glyQS antitermination: a new twist on tRNA. RNA. 2003;9:1148–1156. doi: 10.1261/rna.5540203. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.McDaniel BA, Grundy FJ, Henkin TM. A tertiary structural element in S box leader RNAs is required for S-adenosylmethionine-directed transcription termination. Mol. Microbiol. 2005;57:1008–1021. doi: 10.1111/j.1365-2958.2005.04740.x. [DOI] [PubMed] [Google Scholar]
  • 51.Grundy FJ, Moir TR, Haldeman MT, Henkin TM. Sequence requirements for terminators and antiterminators in the T box transcription antitermination system: disparity between conservation and functional requirements. Nucleic Acids Res. 2002;30:1646–1655. doi: 10.1093/nar/30.7.1646. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Grundy FJ, Henkin TM. tRNA as a positive regulator of transcription antitermination in B. subtilis. Cell. 1993;74:475–482. doi: 10.1016/0092-8674(93)80049-k. [DOI] [PubMed] [Google Scholar]
  • 53.Anagnostopoulos C, Spizizen J. Requirements for transformation in Bacillus subtilis. J. Bacteriol. 1961;81:741–746. doi: 10.1128/jb.81.5.741-746.1961. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Miller JH. Experiments in Molecular Genetics. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press; 1972. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

supplmental_data_all

RESOURCES