Selective mRNA Degradation by Polynucleotide Phosphorylase in ...

3 downloads 0 Views 2MB Size Report
Upon cold shock, Escherichia coli cell growth transiently stops. During this ... acclimation phase are essential for cellular adaptation to low temperature (55, 60 ...
JOURNAL OF BACTERIOLOGY, May 2001, p. 2808–2816 0021-9193/01/$04.00⫹0 DOI: 10.1128/JB.183.9.2808–2816.2001 Copyright © 2001, American Society for Microbiology. All Rights Reserved.

Vol. 183, No. 9

Selective mRNA Degradation by Polynucleotide Phosphorylase in Cold Shock Adaptation in Escherichia coli KUNITOSHI YAMANAKA†

AND

MASAYORI INOUYE*

Department of Biochemistry, Robert Wood Johnson Medical School, Piscataway, New Jersey 08854 Received 9 August 2000/Accepted 20 February 2001

Upon cold shock, Escherichia coli cell growth transiently stops. During this acclimation phase, specific cold shock proteins (CSPs) are highly induced. At the end of the acclimation phase, their synthesis is reduced to new basal levels, while the non-cold shock protein synthesis is resumed, resulting in cell growth reinitiation. Here, we report that polynucleotide phosphorylase (PNPase) is required to repress CSP production at the end of the acclimation phase. A pnp mutant, upon cold shock, maintained a high level of CSPs even after 24 h. PNPase was found to be essential for selective degradation of CSP mRNAs at 15°C. In a poly(A) polymerase mutant and a CsdA RNA helicase mutant, CSP expression upon cold shock was significantly prolonged, indicating that PNPase in concert with poly(A) polymerase and CsdA RNA helicase plays a critical role in cold shock adaptation.

E was shown to be involved in the extreme instability of cspA mRNA at high temperatures (19). The 5⬘-UTR of the cspA mRNA contains a unique sequence called the cold box (29). When fragments containing the coldbox sequence were overproduced at 15°C, CspA production became constitutive. However, when CspA was simultaneously overproduced, the normal transient expression was resumed (29). Moreover, when a cspA gene without the cold-box sequence was reintroduced into a cspA deletion mutant, CspA production became constitutive due to poor repression at the end of the acclimation phase (2, 18). These results indicated that the cold-box sequence plays an important role in autoregulation of the cspA expression. In addition, mRNA stability was also suggested to play a role in the repression of cspA expression in the repression stage of the acclimation phase on the basis that the half-life of the cspA mRNA became shorter (20). The facts that mRNA stability is the major factor for regulation of CSP expression (6, 19, 20) and that polynucleotide phosphorylase (PNPase) is a cold shock-inducible exoribonuclease and a component of RNA degradosomes involved in mRNA degradation (1, 3, 4, 16, 33, 43, 49, 55, 60) led us to examine whether PNPase is involved in the regulation of CSP expression upon cold shock. Here, we report that PNPase is essential for the adaptive growth resumption of cold shocktreated cells by selectively degrading mRNAs for CSPs at the end of the acclimation phase. In a pnp mutant, the induction of CSPs upon cold shock was normally observed as in the wildtype strain. However, their production no longer autoregulated and was maintained at high levels throughout cold shock treatment. This resulted in cell growth arrest and a dramatic reduction of colony-forming ability below 25°C. Importantly, during preparation of this paper it was reported that a PNPase-deficient mutant of Yersinia enterocolitica was unable to degrade cspA1/A2 mRNA properly after cold shock (46). Taken together, these results indicate that the PNPase function is required for the critical transition from the acclimation phase to cell growth resumption after cold shock.

Upon temperature downshift, Escherichia coli cells rapidly but transiently produce a selective set of proteins called cold shock proteins (CSPs), which are considered to be essential for cellular adaptation to low temperature (33, 60). At 15°C the synthesis of CSPs dramatically increases during the first hour (induction stage) and then is reduced to new basal levels (repression stage). During this period (acclimation phase), including both the induction and repression stages, cell growth is arrested, indicating that cellular events occurring during the acclimation phase are essential for cellular adaptation to low temperature (55, 60, 61). Unlike the heat shock response, in which a heat shock sigma factor (␴32) plays a major role in the induction of heat shock proteins (66), a specific sigma factor for cold shock response has not been identified. It has been shown that no de novo protein synthesis is required for the induction of CSPs (17). CspA is considered to be a major cold shock protein (22, 63), and its expression is regulated in a complex manner at the levels of transcription, mRNA stability, and translation efficiency (60). Among these, mRNA stability is considered to play a major role in cold shock induction (6, 19, 20). Importantly, the cspA mRNA possesses an unusually long 5⬘ untranslated region (5⬘-UTR) consisting of 159 bases (54), which is crucial for the mRNA stability (6, 19, 20, 64). Although cspA is well transcribed at 37°C (21, 44), CspA is almost undetectable at this temperature due to extreme instability of the cspA mRNA except for immediately after dilution of a stationaryphase culture into a fresh rich medium and upon nutritional upshift (7, 63). The cspA mRNA is dramatically stabilized upon cold shock, allowing a high CspA production (6, 19, 20). RNase

* Corresponding author. Mailing address: Department of Biochemistry, Robert Wood Johnson Medical School, 675 Hoes La., Piscataway, NJ 08854. Phone: (732) 235-4115. Fax: (732) 235-4559. E-mail: [email protected]. † Present address: Division of Molecular Cell Biology, Institute of Molecular Embryology and Genetics, Kumamoto University, 4-24-1 Kuhonji, Kumamoto, 862-0976 Japan. 2808

VOL. 183, 2001

PNPase AND COLD SHOCK ADAPTATION

2809

FIG. 1. Two-dimensional analysis of total cellular proteins upon cold shock in pnp⫹ and pnp-7 cells. Cells were labeled with [35S]methionine at 37°C (a and e) and then at 0.5 h (b and f), 4 h (c and g), and 20 h (d and h) after temperature downshift from 37 to 15°C. Cell lysates were analyzed by two-dimensional gel electrophoresis as described previously (31). The first dimension was carried out within the range of 3.5 (right) and 10 (left). (a to d) The wild-type strain MG1693. (e to h) The pnp-7 mutant SK5691. Typical CSPs are circled. Spot no. 1, PNPase (33); no. 2, CsdA (32); no. 3, RbfA (31); no. 4, CspA and CspG (53); no. 5, CspI (53); and no. 6, CspB (53).

MATERIALS AND METHODS Bacterial strains. E. coli strains MG1693 (thyA715) (1), SK5691 (thyA715 pnp-7) (1), SH3208 (his ⌬trpE5 ␭) (35), BZ452 (his ⌬trpE5 smbB105 zcb::Tn10 ␭) (35), MC4100 [F⫺ araD139 ⌬(argF-lac)U169 rpsL150 relA1 flbB5301 deoC1 ptsF25 rbsR] and AR134 (MC4100 but pcnB80) (25) were used in this study. pnp-7 appeared to be a nonsense mutation (1). smbB is an allele of rne, and smbB105 also possesses a nonsense mutation to produce C-terminal half-truncated RNase E (35). MG1693 and SK5691 were provided by S. R. Kushner (University of Georgia), and SH3208 and BZ452 were provided by S. Hiraga (Kumamoto University, Japan). Construction of a csdA null mutant. The kanamycin-resistant gene (1.3-kb HincII fragment) from pUC7Km(Pst) was inserted into the middle of the coding region of the csdA gene (alternatively, deaD and mssB) at the EcoRV site in pKX164 (65), yielding pKNJ9026. The linearized pKNJ9026 DNA fragment containing the disrupted csdA gene (csdA::kan) was then introduced into the chromosome of a recD mutant FS1576 (recD1009 thi-1 thr-1 leuB6 lacY1 tonA21 supE44). Kanamycin-resistant transformants were isolated, and the disruption of csdA on the chromosome was confirmed by Southern hybridization (data not shown). The csdA::kan mutation was transduced into the wild-type strain MC4100, yielding KNJ130 (MC4100 but csdA::kan). Protein labeling experiments. Cells were grown in M9 medium supplemented with glucose, 19 amino acids (no methionine), and thymine at 37°C and then were transferred to 15°C. Cells were labeled with [35S]methionine (1,092 Ci/ mmol; Amersham) for 10 min. For a functional stability assay of mRNAs, rifampin was added to a final concentration of 200 ␮g/ml at 1 h after transfer to 15°C and then cells were labeled with [35S]methionine for 5 min at 0, 20, 40, and 80 min after the addition of rifampin. Cell lysates were prepared and processed by two-dimensional gel electrophoresis as described previously (31). The first dimension was carried out within the pH range of 3.5 (right) and 10 (left). Western blot analysis. Cells were grown in M9 medium supplemented with glucose, Casamino Acids, L-tryptophan, and thymine at 37°C and then were

transferred to 15°C. Cells were collected at 0, 4, and 20 h after the temperature downshift, and colony-forming abilities were examined by using Luria-Bertani (LB) plates containing thymine at 37°C. Equal numbers of cells (107 cells) were analyzed by a tricine–sodium dodecyl sulfate (SDS)–16.5% polyacrylamide gel (52) for CspA or by an SDS–12.5% polyacrylamide gel for Era and enolase. After transfer to a nitrocellulose membrane (BA85), CspA, Era, and enolase were detected by using anti-CspA antiserum (54), anti-Era antiserum (42), and antienolase antiserum (gift from A. Carpousis), respectively, as described previously (63). Alkaline phosphatase-conjugated goat anti-rabbit immunoglobulin G (IgG; Sigma) was used as a second antibody. Detection was carried out by using the chromogenic substrates 5-bromo-4-chloro-3-indolyl-phosphate (BCIP) and 4-nitroblue tetrazolium chloride (NBT) (Boehringer Mannheim). Isolation of RNA and primer extension analysis. Cells were grown in M9 medium supplemented with glucose, Casamino Acids, L-tryptophan, and thymine at 37°C and then were transferred to 15°C. Cells were collected at 0, 0.5, and 4 h after the temperature downshift, and RNA was extracted by the hot-phenol method as described previously (62). Primer no. 8285, 5⬘-ACATAGTGTATTA CCTTTAA-3⬘, which corresponds to the complementary strand of ⫹144 to ⫹163 of cspA, where the transcription start site was assigned as ⫹1 (54), was labeled with [␥-32P]ATP (⬎5,000 Ci/mmol; DuPont-New England Nuclear) by T4 polynucleotide kinase (Gibco BRL). Primer extension was carried out using avian myeloblastosis virus reverse transcriptase (Boehringer Mannheim) with 5 ␮g of RNA as described previously (62). Primer extension products were analyzed on a 6% polyacrylamide gel under denaturing conditions in 1⫻ Tris-borateEDTA (TBE) and 6 M urea.

RESULTS Effect of a pnp mutation on protein expression upon cold shock. In the wild-type strain, CSPs, such as CspA, RbfA, and CsdA, were almost undetectable at 37°C (Fig. 1a) but were

2810

YAMANAKA AND INOUYE

FIG. 2. Primer extension analysis of the cspA mRNA. RNA extraction and primer extension analysis were carried out as described in Materials and Methods. The primer extension products were analyzed on a 6% denatured polyacrylamide gel. Lanes 1, 4, 7, and 10, RNA extracted from exponentially growing cells at 37°C; lanes 2, 5, 8, and 11, RNA extracted from cells at 0.5 h after cold shock; lanes 3, 6, 9, and 12, RNA extracted from cells at 4 h after cold shock. Lanes 1 to 3, MG1693 (pnp⫹); lanes 4 to 6, SK5691 (pnp-7); lanes 7 to 9, SH3208 (rne⫹); lanes 10 to 12, BZ452 (smbB105). The relative cspA mRNA amounts were calculated using the transcript at 0.5 h of each strain as 100%.

dramatically induced upon temperature downshift to 15°C (Fig. 1b). At 4 h (Fig. 1c) and 20 h (Fig. 1d) after the downshift, the expression of individual CSPs in the wild-type cells was reduced to a new basal level. In contrast to CSP expression, the synthesis of most of the non-cold shock proteins (non-CSPs) was transiently inhibited during the acclimation phase. However, after the acclimation phase their synthesis was resumed with concomitant repression of the CSP expression (see Fig. 1c). This is consistent with the notion that cell growth is resumed approximately 2 h after cold shock to 15°C (22). Next, the CSP expression was examined in a pnp mutant. As shown in Fig. 1e, CSPs were hardly detected at 37°C in the pnp mutant, as in the wild-type strain (Fig. 1a), suggesting that PNPase is not essential for the degradation of at least the cspA mRNA at 37°C and that the cspA mRNA is extremely unstable at 37°C in both strains. However, it should be mentioned that there are significant differences between the wild-type strain and the pnp mutant at 37°C (compare Fig. 1a with 1e), suggesting that the loss of PNPase affects gene expression probably through mRNA degradation. Upon temperature downshift, CSPs in the pnp mutant were dramatically induced, while the synthesis of other cellular proteins was significantly reduced (Fig. 1f) in a manner similar to that of the wild-type strain (Fig. 1b). This indicates that the pnp mutation does not affect the initial cellular response to temperature downshift. However, the synthesis of CSPs in the pnp mutant was no longer transient at 4 h (Fig. 1g). Surprisingly, even at 20 h after temperature downshift, a high level of CSP production was maintained (Fig. 1h). The inability to repress CSP expression in the pnp mutant was confirmed by analyzing the amounts of mRNA. The cspA mRNA analyzed by primer extension was hardly detected at 37°C for the pnp⫹ strain, as shown in Fig. 2, lane 1, whereas it was slightly detected for the pnp-7 strain (Fig. 2, lane 4) (see below). At 0.5 h after temperature downshift it dramatically increased for both pnp⫹ and pnp-7 strains (Fig. 2, lanes 2 and 5, respectively). However, the most significant difference between pnp⫹ and pnp-7 strains can be observed at 4 h after the temperature downshift. In the wild-type strain, the amount of cspA mRNA was reduced to a very low basal level (Fig. 2, lane 3), while in the pnp mutant the level remained unchanged (Fig.

J. BACTERIOL.

FIG. 3. Immunological detection of CspA, Era, and enolase in pnp⫹ and pnp-7 cells after temperature downshift. Cells were grown in M9 medium supplemented with glucose, Casamino Acids, L-tryptophan, and thymine at 37°C to a mid-exponential phase and then were transferred to 15°C. Cells were collected at 0 (lanes 1 and 4), 4 (lanes 2 and 5), and 20 (lanes 3 and 6) h after temperature downshift, and equal numbers of cells (107 cells) per lane were analyzed by a tricine– SDS–16.5% polyacrylamide gel (52) for CspA and by an SDS–12.5% polyacrylamide gel for Era and enolase. After transfer to nitrocellulose membrane (BA85), proteins were detected by using anti-CspA antiserum, anti-Era antiserum, or anti-enolase antiserum. Alkaline phosphatase-conjugated goat anti-rabbit IgG (Sigma) was used as a second antibody. Detection was carried out by using the chromogenic substrates BCIP and NBT. Lanes 1 to 3, MG1693 (pnp⫹); lanes 4 to 6, SK5691 (pnp-7).

2, lane 6), where similar amounts of the cspA mRNA were detected at both 0.5 and 4 h after temperature downshift (compare Fig. 2, lanes 6 and 5). The apparent derepression of the cspA mRNA production was further confirmed by Western blot analysis as shown in Fig. 3. In comparison with the wild-type strain, the amount of the cold shock-induced CspA in the pnp mutant was higher (approximately fourfold) at 4 h (compare lane 5 with lane 2) and extraordinarily higher (⬎10-fold) at 20 h (compare lane 6 with lane 3), indicating that CspA production was not repressed in the pnp mutant. Although the cspA mRNA was slightly detected at 37°C for the pnp mutant as described above (Fig. 2, lane 4), CspA protein was not detected by Western blot analysis (Fig. 3, lane 4). This is consistent with the previous notion that translational control as well as mRNA stability plays an important role in cspA expression at 37°C (19, 64). It should be noted that the anti-CspA antiserum used was highly specific against CspA, as it does not interact with any other CspA homologues (CspB to CspI) having 44 to 80% identity to CspA (58, 61, 63). Most importantly, the unusual accumulation observed for CspA in the pnp mutant cannot be observed for other non-CSPs such as Era and enolase (Fig. 3); their amounts were almost identical in both pnp⫹ and pnp-7 strains at either 37 or 15°C, suggesting that PNPase may be involved in the expression regulation of a selective set of proteins at 15°C. Selective degradation of mRNAs by PNPase. The results described above indicate that the cspA expression in the pnp-7 mutant is normally induced upon cold shock but is not autoregulated even in the presence of a very high concentration of CspA. In the wild-type strain, CspA production upon cold shock is transient and the rate of CspA synthesis is rapidly

VOL. 183, 2001

PNPase AND COLD SHOCK ADAPTATION

2811

FIG. 4. Analysis of the functional stability of mRNA. Cells were grown at 37°C to mid-exponential phase, transferred to 15°C, and then further incubated for 1 h. Rifampin was added to a final concentration of 200 ␮g/ml. Cells were labeled with [35S]methionine for 5 min at 0 min (a and e), 20 min (b and f), 40 min (c and g) and 80 min (d and h) after the addition of rifampin. Cell lysates were analyzed by two-dimensional gel electrophoresis under the same conditions as in Fig. 1. (a to d) The wild-type strain MG1693. (e to h) The pnp-7 mutant SK5691. Typical CSPs are circled. Spot no. 1, PNPase; no. 2, CsdA; no. 3, RbfA; no. 4, CspA and CspG; no. 5, CspI; no. 6, CspB. Forty spots were densitometrically measured, and functional half-lives were calculated from two experiments.

reduced in the later half of the acclimation phase (22). Since the change in the rate of CspA synthesis has been shown to be parallel to the change in the amount of cspA mRNA, the cspA autoregulation during the acclimation phase is considered to be at the level of mRNA. Therefore, we next attempted to analyze the mRNA stability not only of cspA but also of other genes for CSPs and nonCSPs by measuring the functional half-lives of their mRNAs in both pnp⫹ and pnp-7 strains at the middle of the acclimation phase. At 1 h after temperature downshift from 37 to 15°C, rifampin was added to a final concentration of 200 ␮g/ml to inhibit transcription initiation. Cellular proteins were then labeled for 5 min with [35S]methionine at 0, 20, 40, and 80 min after the addition of rifampin. In the wild-type strain, almost all cellular mRNAs, including mRNAs for CSPs, were stable at least for the first 20 min (compare Fig. 4b to 4a). However, expression of CSPs greatly decreased at 40 min after the addition of rifampin together with other non-CSPs (Fig. 4c). At 80 min, CSP production was dramatically reduced, while several non-CSPs were still synthesized (Fig. 4d). In contrast, CSP synthesis in the pnp mutant was almost unaffected until 40 min after the addition of rifampin (Fig. 4e to 4g). Even at 80 min, CSPs were still produced at significantly high levels (Fig. 4h). Subsequently, functional half-lives of mRNAs encoding a number of proteins were determined densitometrically. Interestingly, functional half-lives of mRNAs for 24 out of 27 nonCSPs analyzed in the pnp mutant were found to be shorter than

those in the wild-type strain (see Discussion). In contrast, mRNAs for CSPs in the pnp mutant became more stable than those in the wild-type strain. For example, the half-life of the mRNA for cspA and cspG in the pnp mutant was approximately 60 min, while that in the wild-type strain was 30 min (spot 4 in Fig. 4; note that CspA and CspG were inseparable in the gels used). This is consistent with the previous result that the functional half-life of the cspA mRNA in the wild-type strain at 1 h after temperature downshift from 37 to 15°C is about 30 min (54). Similarly, the functional half-lives of mRNAs for CspI (spot 5) and CspB (spot 6) increased from 31 and 24 min, respectively, in the wild-type strain to 38 and 54 min, respectively, in the pnp mutant. The half-lives of mRNAs for five out of seven other CSPs analyzed changed from 8, 17, 24, 38, and 68 min in the wild-type strain to 10, 40, 37, 44, and ⬎80 min, respectively, in the pnp mutant, although those for two CSPs (CsdA [spot 2] and RbfA [spot 3]) were slightly reduced (from 17 and 34 min, respectively, in the wild-type strain to 13 and 30 min, respectively, in the pnp mutant). Requirement of PNPase for cell growth at low temperature. Since the synthesis of CSPs was not repressed in the pnp mutant (Fig. 1 and 2), it is speculated that the pnp mutant may not be able to adapt to low temperature, resulting in a cell growth defect. As shown in Fig. 5, the pnp mutant was extremely sensitive to low temperature and was unable to form colonies below 25°C, which agrees well with the results reported by Luttinger et al. (41); the colony-forming ability

2812

YAMANAKA AND INOUYE

dropped to 10⫺7. The cold-sensitive growth phenotype of the pnp mutant was fully complemented by transforming cells with plasmid pKX150 carrying the pnp⫹ gene (65), indicating that PNPase function is essential for growth at low temperature but not at high temperature. Effect of C-terminal half-truncated RNase E on CSP expression. PNPase is known to interact with the C-terminal region of RNase E and to form the RNA degradosome in association with RhlB (an RNA helicase), DnaK, and enolase (43, 49, 57). Recently it was reported that only RNase E, PNPase, and RhlB are enough to reconstitute a functional RNA degradosome (13). At low temperature, more stable stem-loop structures are formed in mRNAs, which makes them more resistant to degradation by exoribonucleases, such as PNPase. Among the RNA degradosome components, only PNPase is known to be cold shock inducible. Therefore, it is interesting to determine whether cold shock-induced PNPase constitutes a functional component of the RNA degradosome. To test this question, we used the smbB105 mutant, in which the C-terminal half of RNase E is truncated (35), resulting in no formation of the RNA degradosome (35, 52). As shown in Fig. 6, CSPs were not detected at 37°C in either wild-type or smbB105 strains (Fig. 6a and 6d, respectively), while they were dramatically induced immediately upon cold shock in both strains (Fig. 6b and 6e, respectively). At 4 h after temperature downshift, their expression in the mutant was repressed as in the wild-type strain (Fig. 6f and 6c, respectively). The changes in the amounts of cspA mRNA in the smbB105 mutant upon cold shock (Fig. 2, lanes 10 to 12) were also very similar to those observed in the wild-type strain (Fig. 2, lanes 7 to 9). Furthermore, cell growth was not affected by the smbB105 mutation at either high or low temperatures (data not shown). These results indicate that the association of PNPase with RNase E or the degradosome formation is not required for the adaptive degradation of CSP mRNAs by PNPase. Effect of a pcnB mutation on CSP expression. It is reasonable to assume that the key for the selective degradation of a group of mRNAs exists in the 3⬘ end of mRNAs, since PNPase is a 3⬘ to 5⬘ exoribonuclease. Not only in eukaryotes but also in prokaryotes, polyadenylation is a well-known, posttranscriptional modification at the 3⬘ end of mRNAs (51). E. coli contains poly(A) polymerase encoded by pcnB (9), and poly(A) tail has been suggested to regulate mRNA degradation (5, 11, 12, 24, 26, 28, 47). Moreover, purified PNPase preferentially degrades polyadenylated RNA but not nonpolyadenylated RNA when these two species exist simultaneously (39). If CSP mRNAs are selectively polyadenylated, one can expect that CSP synthesis should be prolonged in a pcnB mutant due to reduced degradation. To examine if poly(A) tail is involved in CSP expression, we carried out the pulse-labeling experiment using the pcnB80 mutant and subsequently Western blotting analysis with the anti-CspA antibody (25, 40). Interestingly, CspA expression was clearly detected in the pcnB80 mutant at 37°C (Fig. 7e), while that in the wild-type strain was hardly detected (Fig. 7a). Indeed, the amount of CspA in the pcnB80 mutant was a little higher than that in the wild-type strain (data not shown). These results suggest that polyadenylation is involved in CspA expression to some extent; that is, polyadenylation might enhance the degradation of CspA mRNA. Upon temperature downshift, CSPs were dramatically induced in the

J. BACTERIOL.

FIG. 5. Colony-forming abilities of the pnp mutant at various temperatures. Cultures of strains were grown in LB medium containing thymine (50 ␮g/ml) and chloramphenicol (30 ␮g/ml) at 37°C overnight. After appropriate dilutions, cells were plated on LB agar plates containing thymine (50 ␮g/ml) and chloramphenicol (30 ␮g/ml) and were incubated at 42, 37, 30, 25, 20, and 15°C. , the wild-type strain MG1693 carrying the vector pHSG575; ‚, MG1693 carrying the pnp⫹ plasmid pKX150 (65); , the pnp-7 mutant SK5691 carrying pHSG575; Œ, SK5691 carrying pKX150.

wild-type cells (Fig. 7b) as well as in the pcnB80 mutant cells (Fig. 7f). However, importantly the synthesis of CSPs in the pcnB80 mutant was prolonged to 4 h after temperature downshift (compare Fig. 7g to 7c), and at 20 h it dropped to a new basal level, which is a level similar to that in the wild-type strain (Fig. 7d and 7h). Consistently, it was found that the amount of CspA in the pcnB80 mutant was significantly higher at 20 h after temperature downshift than that in the wild-type strain (data not shown). In contrast, the amount of enolase, a nonCSP, was unchanged upon temperature downshift (data not shown). These results indicate that polyadenylation plays an important role in the degradation of CSP mRNAs. It should be mentioned, however, that in contrast to PNPase, the pcnB80 mutant did not show the cold-sensitive growth phenotype (data not shown) and poly(A) polymerase was not a CSP from the analysis of pcnB::lacZ fusion constructs (data not shown). Effect of a csdA mutation on CSP expression. It is interesting to note that poly(A) polymerase of E. coli was reported to interact with RNA, RNase E, and DEAD-box RNA helicases of RhlE, SrmB, and CsdA by means of far Western analysis (50). Among them, CsdA is a cold shock-inducible protein and has an activity to unwind double-stranded RNA in the absence of ATP (32). CsdA function was proposed to be essential for ribosome function to increase translational efficiencies of mRNAs by unwinding stable secondary structures formed at low temperature (32). To examine whether CsdA is involved in

VOL. 183, 2001

PNPase AND COLD SHOCK ADAPTATION

2813

FIG. 6. Two-dimensional analysis of total cellular proteins upon cold shock in rne⫹ and smbB105 cells. Cells were labeled with [35S]methionine at 37°C (a and d) and then at 0.5 h (b and e) and 4 h (c and f) after temperature downshift from 37 to 15°C. Cell lysates were analyzed by two-dimensional gel electrophoresis under the same conditions as in Fig. 1. (a to c) The wild-type strain SH3208. (d to f) The smbB105 mutant BZ452. Typical CSPs are circled in b and e. Spot no. 1, PNPase; no. 2, CsdA; no. 3, RbfA; no. 4, CspA and CspG; no. 5, CspI; no. 6, CspB.

the selective degradation of CSP mRNAs after the acclimation phase, we prepared a csdA null mutant as described in Materials and Methods. CSPs were hardly detected in the csdA mutant at 37°C (Fig. 7i), as in the wild-type strain (Fig. 7a). Consistently, the amount of CspA in the csdA mutant at 37°C was almost undetectable (data not shown). Upon temperature downshift to 15°C, CSPs were dramatically induced in the csdA mutant (Fig. 7j). Similar to the pcnB mutant described above, CSP expression in the csdA mutant was prolonged to 4 h after temperature downshift (Fig. 7k) and it was dropped to the new basal level at 20 h (Fig. 7l). The amount of CspA in the csdA mutant was consistent with these results (data not shown). The amount of enolase was again not affected by the csdA mutation (data not shown). It should also be mentioned that the csdA mutant did grow at low temperature with a reduced growth rate (data not shown). DISCUSSION The present work demonstrates that PNPase is involved in selective degradation of CSP mRNAs at the stage of repression during the acclimation phase. Very recently it was reported that a PNPase-deficient mutant of Y. enterocolitica was unable to degrade cspA1/A2 mRNA properly after cold shock (46). The authors suggested that after synthesis of CSPs and cold

adaptation of the cells, CSP mRNAs must be degraded; otherwise, they trap ribosomes, prevent translation of bulk mRNAs, and thus inhibit growth of the bacterium at low temperatures (46). Our results demonstrated here are fully consistent with their observations, and these results together suggest that PNPase regulates gene expression by selectively degrading specific mRNAs. Such a loss of autoregulation of CSP genes during the process of cold shock adaptation has also been observed in another CSP mutant, an rbfA mutant. rbfA encodes a cold shock protein required for ribosomal maturation and/or translation initiation. A constitutive induction of the cold shock response occurs in this mutant in a manner similar to that found in the present study, resulting in slower cell growth (31). It is thus apparent that continuous high expression of CSPs is deleterious to cells and that for cell growth resumption, their expression has to be repressed once the cellular concentrations of individual CSPs reach optimal levels. PNPase was reported to be cold shock inducible in E. coli (33), Y. enterocolitica, a psychrotrophic bacterium (23), and Photorhabdus sp. (family Enterobacteriaceae) (10). In E. coli grown at 37°C, PNPase was shown to be responsible for only 10% of the total processive 3⬘ to 5⬘ mRNA degradation (15). It is important to note that the phosphorolytic degradation of mRNA by PNPase is less energetically expensive than the hydrolytic degradation by RNase II, another exoribonuclease

2814

YAMANAKA AND INOUYE

J. BACTERIOL.

FIG. 7. Two-dimensional analysis of total cellular proteins upon cold shock in pcnB80 and csdA mutant cells. Cells were labeled with [35S]methionine at 37°C (a, e, and i) and then at 0.5 h (b, f, and j), 4 h (c, g, and k), and 20 h (d, h, and l) after temperature downshift from 37 to 15°C. Cell lysates were analyzed by two-dimensional gel electrophoresis under the same conditions as in Fig. 1. Only a portion corresponding to low-molecular-weight proteins is shown. (a to d) The wild-type strain MC4100. (e to h) The pcnB80 mutant AR134. (i to l) The csdA::kan mutant KNJ130. Typical CSPs are indicated by an arrow. Spot no. 1, RbfA; no. 2, CspA and CspG; no. 3, CspI; no. 4, CspB.

in E. coli, as the high-energy phosphate bonds of the ribonucleotide are retained in the PNPase reaction (14, 15). Thus, PNPase may be more favorable than RNase II for mRNA degradation in E. coli at low temperature. In a pnp mutant of Y. enterocolitica, cell growth was severely restricted at 5°C but not at 30°C (23), and in the case of Bacillus subtilis, a pnp mutation caused cold-sensitive cell growth (41, 59). Luttinger et al. (41) and we also (in Fig. 5) demonstrated that the E. coli pnp mutant showed a cold-sensitive growth phenotype. All these results strongly indicate the importance of PNPase under the low-temperature growth condition. It should be noted that in B. subtilis, PNPase has been suggested to be essential for the expression of specific genes at the posttranscriptional level, mRNA stability, or translation (41). An important feature of PNPase is that it contains two RNA-binding domains, a KH domain and an S1 domain (8). The S1 domain was originally identified in ribosomal protein S1 (53). The three-dimensional structure of the S1 domain of PNPase has been determined and found to contain a ␤-barrel structure similar to that of CspA (8). CspA is an RNA-binding protein and has been proposed to function as an RNA chaperone (30). Furthermore, PNPase (39) and ribosome protein S1 (34) were reported to preferentially bind to the poly(A) sequence, and PNPase preferably degrades polyadenylated RNA (39). In plant chloroplasts, it has also been shown that PNPase preferentially degrades polyadenylated RNA (38) and that PNPase is a component of the poly(A) polymerase complex (36). Very recently, Mohanty and Kushner reported that PNPase functions as both an exonuclease and a poly(A) polymerase in E. coli (45). While there are few sequence homologies at the 3⬘-UTRs among cspA, cspB, cspG, and cspI, the addition of poly(A) tail to CSP mRNAs may result in selective

degradation of these mRNAs by PNPase at low temperature (see Fig. 7). Therefore, at least at low temperature, PNPase in concert with poly(A) polymerase is reasonably considered to play an important role in the selective degradation of CSP mRNAs. However, it is also possible that an additional unknown factor(s) may selectively recognize mRNAs for CSPs to enhance their degradation by PNPase or that PNPase may selectively recognize a common sequence such as the cold box in the 5⬘-UTR of CSP mRNAs to enhance their degradation. In any case, the precise mechanism of selective degradation of CSP mRNAs by PNPase at low temperature awaits further characterization. PNPase is known to form the RNA degradosome together with at least RNase E and RhlB (13, 57). The endoribonuclease RNase E and the RNA helicase RhlB are speculated to cooperatively act with PNPase for mRNA degradation. In the present study, however, we demonstrated that the interaction of PNPase with RNase E and RhlB is not required for the function of PNPase and cell growth at low temperature (see Fig. 2 and 6). Therefore, the RNA degradosomes are unlikely to play major roles in cells growing at low temperature. Very recently it was reported that although RNA degradosomes indeed exist in vivo in E. coli as multicomponent complexes, PNPase and enolase are present in E. coli in large excess relative to RNase E (37). Therefore, PNPase is detected in cells largely as molecules unlinked to the RNase E (37). Furthermore, the assembly of the RNase E-based degradosome of E. coli was reported not to be required for normal mRNA decay in vivo (48). Our results presented here are fully consistent with these results. It should be mentioned that poly(A) polymerase was also shown to interact with RNA, RNase E, and DEAD-box RNA helicases (RhlE, SrmB, and CsdA) (50),

VOL. 183, 2001

PNPase AND COLD SHOCK ADAPTATION

among which CsdA is cold shock inducible (32). Indeed, in the csdA mutant used in this study, CSP expression was significantly prolonged (see Fig. 7), suggesting that CsdA may play a role in unwinding RNA structures that impede the processive activity of PNPase. Note that the CsdA unwinding activity is independent of ATP (32), which might be an advantage in terms of energy consumption at low temperature. Taken together, PNPase, poly(A) polymerase, and CsdA are involved in the efficient and selective degradation of CSP mRNAs, which is essential for the cold shock adaptation of E. coli. In the pnp mutant used in the present work, CSP mRNAs were specifically stabilized approximately twofold (Fig. 4), suggesting that PNPase is associated with the degradation of CSP mRNAs. Interestingly, in contrast to CSP mRNAs, non-CSP mRNAs became more unstable in the pnp mutant than in the wild-type strain. It is tempting to speculate that increased concentrations of CspA and other CspA homologues in the mutant cells destabilize non-CSP mRNAs as CspA and its homologues function as RNA chaperones (30). In conclusion, the present results together with a recent report (46) represent that a specific ribonuclease plays an essential role in stress adaptation by selectively degrading mRNAs for stress-response proteins. As the overproduction of stress-response proteins appears to be deleterious to cells, the overproduction is prevented by a ribonuclease, which is induced by the same stress. In the case of the heat shock response, ␴32 plays the major role (66) and it has been shown that ␴32 is degraded by a specific protease, FtsH, which is also induced by heat shock, to regulate the heat shock response (27, 56). Thus, it is important to mention that a specific ribonuclease plays a critical role in cold shock response and adaptation in contrast to heat shock response and adaptation, where a specific protease plays a major role. This is consistent with the fact that the fate of individual mRNAs for each CSP plays a central role in cold shock. Cold shock stress is likely to be a major stress for most of the prokaryotes in nature, and PNPase may be essential for low-temperature survival and proliferation of not only E. coli and B. subtilis but also many other prokaryotes. ACKNOWLEDGMENTS We thank U. Shinde for critical reading of the manuscript, S. R. Kushner and S. Hiraga for strains, and A. Carpousis for anti-enolase antiserum. The present work was supported by a grant from the National Institutes of Health (GM19043). REFERENCES 1. Arraiano, C. M., S. D. Yancey, and S. R. Kushner. 1988. Stabilization of discrete mRNA breakdown products in ams pnp rnb multiple mutants of Escherichia coli K-12. J. Bacteriol. 170:4625–4633. 2. Bae, W., P. G. Jones, and M. Inouye. 1997. CspA, the major cold shock protein of Escherichia coli, negatively regulates its own gene expression. J. Bacteriol. 179:7081–7088. 3. Bae, W., B. Xia, M. Inouye, and K. Severinov. 2000. Escherichia coli CspAfamily RNA chaperones are transcription antiterminators. Proc. Natl. Acad. Sci. USA 97:7784–7789. 4. Beran, R. K., and R. W. Simons. 2001. Cold-temperature induction of Escherichia coli polynucleotide phosphorylase occurs by reversal of its autoregulation. Mol. Microbiol. 39:112–125. 5. Blum, E., A. J. Carpousis, and C. F. Higgins. 1999. Polyadenylation promotes degradation of 3⬘-structured RNA by the Escherichia coli RNA degradosome in vitro. J. Biol. Chem. 274:4009–4016. 6. Brandi, A., P. Pietroni, C. O. Gualerzi, and C. L. Pon. 1996. Post-transcriptional regulation of CspA expression in Escherichia coli. Mol. Microbiol. 19:231–240.

2815

7. Brandi, A., R. Spurio, C. O. Gualerzi, and C. L. Pon. 1999. Massive presence of the Escherichia coli ‘major cold-shock protein’ CspA under non-stress conditions. EMBO J. 18:1653–1659. 8. Bycroft, M., T. J. P. Hubbard, M. Proctor, S. M. V. Freund, and A. G. Murzin. 1997. The solution structure of the S1 RNA binding domain: a member of an ancient nucleic acid-binding fold. Cell 88:235–242. 9. Cao, G.-J., and N. Sarkar. 1992. Identification of the gene for an Escherichia coli poly(A) polymerase. Proc. Natl. Acad. Sci. USA 89:10380–10384. 10. Clarke, D. J., and B. C. A. Dowds. 1994. The gene coding for polynucleotide phosphorylase in Photorhabdus sp. strain K122 is induced at low temperatures. J. Bacteriol. 176:3775–3784. 11. Coburn, G. A., and G. A. Mackie. 1996. Differential sensitivities of proteins of the mRNA for ribosomal protein S20 to 3⬘-exonucleases dependent on oligoadenylation and RNA secondary structure. J. Biol. Chem. 271:15776– 15781. 12. Coburn, G. A., and G. A. Mackie. 1998. Reconstitution of the degradation of the mRNA for ribosomal protein S20 with purified enzymes. J. Mol. Biol. 279:1061–1074. 13. Coburn, G. A., X. Miao, D. J. Briant, and G. A. Mackie. 1999. Reconstitution of a minimal RNA degradosome demonstrates functional coordination between a 3⬘ exonuclease and a DEAD-box RNA helicase. Genes Dev. 13: 2594–2603. 14. Deutscher, M. P. 1993. Ribonuclease multiplicity, diversity and complexity. J. Biol. Chem. 268:13011–13014. 15. Deutscher, M. P., and N. B. Reuven. 1991. Enzymatic basis for hydrolytic versus phosphorolytic mRNA degradation in Escherichia coli and Bacillus subtilis. Proc. Natl. Acad. Sci. USA 88:3277–3280. 16. Donovan, W. P., and S. R. Kushner. 1986. Polynucleotide phosphorylase and ribonuclease II are required for cell viability and mRNA turnover in Escherichia coli K-12. Proc. Natl. Acad. Sci. USA 83:120–124. 17. Etchegaray, J. P., and M. Inouye. 1999. CspA, CspB, and CspG, major cold shock proteins of Escherichia coli, are induced at low temperature under conditions that completely block protein synthesis. J. Bacteriol. 181:1827– 1830. 18. Fang, L., Y. Hou, and M. Inouye. 1998. Role of the cold-box region in the 5⬘ untranslated region of the cspA mRNA in its transient expression at low temperature in Escherichia coli. J. Bacteriol. 180:90–95. 19. Fang, L., W. Jiang, W. Bae, and M. Inouye. 1997. Promoter-independent cold-shock induction of cspA and its derepression at 37°C by mRNA stabilization. Mol. Microbiol. 23:355–364. 20. Goldenberg, D., I. Azar, and A. B. Oppenheim. 1996. Differential mRNA stability of the cspA gene in the cold-shock response of Escherichia coli. Mol. Microbiol. 19:241–248. 21. Goldenberg, D., I. Azar, A. B. Oppenheim, A. Brandi, C. L. Pon, and C. O. Gualerzi. 1997. Role of Escherichia coli cspA promoter sequences and adaptation of translational apparatus in the cold shock response. Mol. Gen. Genet. 256:282–290. 22. Goldstein, J., N. S. Pollitt, and M. Inouye. 1990. Major cold shock proteins of Escherichia coli. Proc. Natl. Acad. Sci. USA 87:283–287. 23. Goverde, R. L. J., J. H. J. Huis in’t Veld, J. G. Kusters, and F. R. Mooi. 1998. The psychrotrophic bacterium Yersinia enterocolitica requires expression of pnp, the gene for polynucleotide phosphorylase, for growth at low temperature (5°C). Mol. Microbiol. 28:555–569. 24. Hajnsdorf, E., F. Braun, J. Haugel-Nielsen, and P. Re´gnier. 1995. Polyadenylation destabilises the rpsO mRNA of Escherichia coli. Proc. Natl. Acad. Sci. USA 90:3973–3977. 25. Harlocker, S. L., A. Rampersaud, W.-P. Yang, and M. Inouye. 1993. Phenotypic revertant mutations of a new OmpR2 mutant (V203Q) of Escherichia coli lie in EnvZ gene, which encodes in the OmpR kinase. J. Bacteriol. 175:1956–1960. 26. Haugel-Nielsen, J., E. Hajnsdorf, and P. Re´gnier. 1996. The rpsO mRNA of Escherichia coli is polyadenylated at multiple sites resulting from endonucleolytic processing and exonucleolytic degradation. EMBO J. 15:3144–3152. 27. Herman, C., D. The´venet, R. D’Ari, and P. Bouloc. 1995. Degradation of ␴32, the heat shock regulation in Escherichia coli, is governed by HflB. Proc. Natl. Acad. Sci. USA 92:3516–3520. 28. Ingle, C. A., and S. R. Kushner. 1996. Development of an in vitro mRNA decay system for Escherichia coli: poly(A) polymerase I is necessary to trigger degradation. Proc. Natl. Acad. Sci. USA 93:12926–12931. 29. Jiang, W., L. Fang, and M. Inouye. 1996. The role of 5⬘-end untranslated region of the mRNA for CspA, the major cold-shock protein of Escherichia coli, in cold-shock adaptation. J. Bacteriol. 178:4919–4925. 30. Jiang, W., Y. Hou, and M. Inouye. 1997. CspA, the major cold-shock protein of Escherichia coli, is an RNA chaperone. J. Biol. Chem. 272:196–202. 31. Jones, P. G., and M. Inouye. 1996. RbfA, 30S ribosomal binding factor, is a cold-shock protein whose absence triggers the cold-shock response. Mol. Microbiol. 21:1207–1218. 32. Jones, P. G., M. Mitta, Y. Kim, W. Jiang, and M. Inouye. 1996. Cold shock induces a major ribosomal-associated protein that unwinds double-stranded RNA in Escherichia coli. Proc. Natl. Acad. Sci. USA 93:76–80. 33. Jones, P. G., R. A. VanBogelen, and F. C. Neidhardt. 1987. Induction of

2816

34. 35. 36. 37.

38. 39. 40. 41. 42. 43. 44.

45. 46. 47. 48. 49.

YAMANAKA AND INOUYE

proteins in response to low temperature in Escherichia coli. J. Bacteriol. 169:2092–2095. Kalapos, M. P., H. Paulus, and N. Sarkar. 1997. Identification of ribosomal protein S1 as a poly(A) binding protein in Escherichia coli. Biochimie 79: 493–502. Kido, M., K. Yamanaka, T. Mitani, H. Niki, T. Ogura, and S. Hiraga. 1996. RNase E polypeptides lacking a carboxyl-terminal half suppress a mukB mutation in Escherichia coli. J. Bacteriol. 178:3917–3925. Li, Q.-S., J. D. Gupta, and A. G. Hunt. 1998. Polynucleotide phosphorylase is a component of a novel plant poly(A) polymerase. J. Biol. Chem. 273: 17539–17543. Liou, G.-G., W.-N. Lane, S. N. Cohen, N.-S. Lin, and S. Lin-Chao. 2001. RNA degradosomes exist in vivo in Escherichia coli as multicomponent complexes associated with the cytoplasmic membrane via the N-terminal region of ribonuclease E. Proc. Natl. Acad. Sci. USA 98:63–68. Lisitsky, I., A. Kolter, and G. Schuster. 1997. The mechanism of preferential degradation of polyadenylated RNA in the chloroplast. J. Biol. Chem. 272: 17648–17653. Lisitsky, I., and G. Schuster. 1999. Preferential degradation of polyadenylated and polyuridinylated RNAs by the bacterial exoribonuclease polynucleotide phosphorylase. Eur. J. Biochem. 261:468–474. Lopilato, J., S. Bortner, and J. Beckwith. 1986. Mutations in a new chromosomal gene of Escherichia coli K-12, pcnB, reduced plasmid copy number of pBR322 and its derivatives. Mol. Gen. Genet. 205:285–290. Luttinger, A., J. Hahn, and D. Dubnau. 1996. Polynucleotide phosphorylase is necessary for competence development in Bacillus subtilis. Mol. Microbiol. 19:343–356. March, P., C. G. Lerner, J. Ahnn, X. Cui, and M. Inouye. 1988. The Escherichia coli Ras-like protein (Era) has GTPase activity and is essential for cell growth. Oncogene 2:539–544. Miczak, A., V. R. Kaberdin, C.-L. Wei, and S. Lin-Chao. 1996. Proteins associated with RNase E in a multicomponent ribonucleolytic complex. Proc. Natl. Acad. Sci. USA 93:3865–3869. Mitta, M., L. Fang, and M. Inouye. 1997. Deletion analysis of cspA of Escherichia coli: requirement of the AT-rich UP element for cspA transcription and the downstream box in the coding region for its cold shock induction. Mol. Microbiol. 26:321–335. Mohanty, B. K., and S. R. Kushner. 2000. Polynucleotide phosphorylase functions both as a 3⬘35⬘ exonuclease and a poly(A) polymerase in Escherichia coli. Proc. Natl. Acad. Sci. USA 97:11966–11971. Neuhaus, K., S. Rapposch, K. P. Francis, and S. Scherer. 2000. Restart of exponential growth of cold-shocked Yersinia enterocolitica occurs after downregulation of cspA1/A2 mRNA. J. Bacteriol. 182:3285–3288. O’Hara, E. E., J. J. Chekavora, C. A. Ingle, Z. R. Kushner, E. Peters, and S. R. Kushner. 1995. Polyadenylation helps regulate mRNA decay in Escherichia coli. Proc. Natl. Acad. Sci. USA 92:1807–1811. Ow, M. C., and S. R. Kushner. 2000. Analysis of mRNA decay and rRNA processing in Escherichia coli in the absence of RNase E-based degradosome assembly. Mol. Microbiol. 38:854–866. Py, B., C. F. Higgins, H. M. Krisch, and A. J. Carpousis. 1996. A DEAD-box

J. BACTERIOL.

50.

51. 52. 53. 54. 55. 56.

57.

58. 59. 60. 61. 62. 63. 64. 65. 66.

RNA helicase in the Escherichia coli RNA degradosome. Nature 381:169– 172. Raynal, L. C., and A. J. Carpousis. 1999. Poly(A) polymerase I of Escherichia coli: characterization of the catalytic domain, an RNA binding site and regions for the interaction with proteins involved in mRNA degradation. Mol. Microbiol. 32:765–775. Sarkar, N. 1997. Polyadenylation of mRNA in prokaryotes. Annu. Rev. Biochem. 66:173–197. Scha ¨gger, H., and G. von Jagow. 1987. Tricine-sodium dodecyl sulfatepolyacrylamide gel electrophoresis for the separation of proteins in the range from 1 to 100 kDa. Anal. Biochem. 166:368–379. Subramanian, A. R. 1983. Structure and functions of ribosomal protein S1. Prog. Nucleic Acid Res. Mol. Biol. 28:101–142. Tanabe, H., J. Goldstein, M. Yang, and M. Inouye. 1992. Identification of the promoter region of the Escherichia coli major cold shock gene, cspA. J. Bacteriol. 174:3867–3873. Thieringer, H. A., P. G. Jones, and M. Inouye. 1998. Cold shock and adaptation. Bioessays 20:49–57. Tomoyasu, T., J. Gamer, B. Bukau, M. Kanemori, H. Mori, A. J. Rutman, A. B. Oppenheim, T. Yura, K. Yamanaka, H. Niki, S. Hiraga, and T. Ogura. 1995. Escherichia coli FtsH is a membrane-bound, ATP-dependent protease which degrades the heat-shock transcription factor ␴32. EMBO J. 14:2551– 2560. Vanzo, N. F., Y. S. Li, B. Py, E. Blum, C. F. Higgins, L. C. Raynal, H. M. Krisch, and A. J. Carpousis. 1998. Ribonuclease E organizes the protein interactions in the Escherichia coli RNA degradosome. Genes Dev. 12:2770– 2781. Wang, N., K. Yamanaka, and M. Inouye. 1999. CspI, the ninth member of the CspA family of Escherichia coli, is induced upon cold shock. J. Bacteriol. 181:1603–1609. Wang, W., and D. H. Bechhofer. 1996. Properties of a Bacillus subtilis polynucleotide phosphorylase deletion strain. J. Bacteriol. 178:2375–2382. Yamanaka, K. 1999. Cold shock response in Escherichia coli. J. Mol. Microbiol. Biotechnol. 1:193–202. Yamanaka, K., L. Fang, and M. Inouye. 1998. The CspA family in Escherichia coli: multiple gene duplication for stress adaptation. Mol. Microbiol. 27:247–255. Yamanaka, K., and M. Inouye. 1997. Growth-phase-dependent expression of cspD, encoding a member of the CspA family in Escherichia coli. J. Bacteriol. 179:5126–5130. Yamanaka, K., and M. Inouye. Induction of CspA, an E. coli major coldshock protein, upon nutritional upshift at 37°C. Genes Cells, in press. Yamanaka, K., M. Mitta, and M. Inouye. 1999. Mutation analysis of the 5⬘ untranslated region of the cold shock cspA mRNA of Escherichia coli. J. Bacteriol. 181:6284–6291. Yamanaka, K., T. Ogura, E. V. Koonin, H. Niki, and S. Hiraga. 1994. Multicopy suppressors, mssA and mssB, of an smbA mutation of Escherichia coli. Mol. Gen. Genet. 243:9–16. Yura, T., H. Nagai, and H. Mori. 1993. Regulation of the heat-shock response in bacteria. Annu. Rev. Microbiol. 47:321–350.