Global Proteomic Analysis of the Insoluble, Soluble, and Supernatant

Dec 1, 2009 - ... Effects of ribosomes and intracellular solutes on activities and stabilities of elongation factor 2 proteins from psychrotolerant an...
0 downloads 0 Views 3MB Size
Global Proteomic Analysis of the Insoluble, Soluble, and Supernatant Fractions of the Psychrophilic Archaeon Methanococcoides burtonii Part I: The Effect of Growth Temperature Timothy J. Williams,†,‡ Dominic W. Burg,†,‡ Mark J. Raftery,§ Anne Poljak,§,| Michael Guilhaus,§ Oliver Pilak,‡ and Ricardo Cavicchioli*,‡ School of Biotechnology and Biomolecular Sciences, The University of New South Wales, Sydney, 2052, NSW, Australia, Bioanalytical Mass Spectrometry Facility, The University of New South Wales, Sydney, 2052, NSW, Australia, and School of Medical Sciences, The University of New South Wales, Sydney, 2052, Australia Received June 10, 2009

The response of the cold-adapted (psychrophilic) methanogenic archaeon Methanococcoides burtonii to growth temperature was investigated using differential proteomics (postincorporation isobaric labeling) and tandem liquid chromatography-mass spectrometry (LC/LC-MS/MS). This is the first proteomic study of M. burtonii to include techniques that specifically enrich for both surface and membrane proteins and to assess the effects of growth temperature (4 vs 23 °C) and carbon source (trimethylamine vs methanol) on cellular protein levels. Numerous surface layer proteins were more abundant at 4 °C, indicating an extensive remodeling of the cell envelope in response to low temperature. Many of these surface proteins contain domains associated with cell adhesion. Within the cell, small proteins each composed of a single TRAM domain were recovered as important cold adaptation proteins and might serve as RNA chaperones, in an analogous manner to Csp proteins (absent from M. burtonii). Other proteins that had higher abundances at 4 °C can be similarly tied to relieving or resolving the adverse affects of cold growth temperature on translational capacity and correct protein folding. The proteome of M. burtonii grown at 23 °C was dominated by oxidative stress proteins, as well as a large number of integral membrane proteins of unknown function. This is the first truly global proteomic study of a psychrophilic archaeon and greatly expands knowledge of the cellular mechanisms underpinning cold adaptation in the Archaea. Keywords: proteome • LC/LC-MS/MS • iTRAQ • membrane protein • archaea • methanogen • psychrophile • methylotroph • cell envelope • S-layer • nucleic acid binding • oxidative stress • translation • transcriptional regulator

Earth’s biosphere is dominated by naturally cold environments (e.g., polar, alpine, deep ocean, seasonally cold habitats), and microbial life has evolved to proliferate throughout all of these regions. Polar environments in particular are very sensitive to changes in global temperature and play critical roles in maintaining microbial processes that are central to ensuring the world’s ecosystems remain balanced and productive.1 Studies examining cold adapted (psychrophilic) members of the Archaea, Bacteria, and Eucarya have revealed that certain molecular mechanisms of adaptation are characteristic of psychrophiles across all lineages.2,3 Some mechanisms of adaptation may be expected to be shared across all psychro-

philes given that microorganisms are isothermal and have to contend with the same impacts that temperatures at or close to freezing have on reaction rates and the structures of cellular components. The latter include impositions caused by increased stability of inhibitory RNA secondary structures, decreased translational capacity, and reduced membrane fluidity.2,4 However, evolution has resulted in some microbial lineages being superior in terms of their capacity to thermally adapt. This is well illustrated by the methanogenic archaea, the only microbial group known to be able to grow at temperatures as high as 122 °C5 or as low as -2 °C6 and all temperatures in between. There is clearly a need to better understand those elements that are fundamental to cold adaptation and those that are specific to particular groups or individual species.

* To whom correpondence should be addressed. Rick Cavicchioli, School of Biotechnology and Biomolecular Sciences, The University of New South Wales, Sydney, 2052, NSW, Australia. E-mail [email protected]; Tel. (+61) 2 9385 3516; Fax (+61) 2 9385 2742. † These authors contributed equally. ‡ School of Biotechnology and Biomolecular Sciences. § Bioanalytical Mass Spectrometry Facility. | School of Medical Sciences.

Within the domain Archaea, few psychrophilic isolates are available for study, and the Antarctic archaeon Methanococcoides burtonii has developed into a robust model for assessing molecular mechanisms of cold adaptation.3 M. burtonii is a methylotrophic methanogen adapted to growth at temperatures of 1-2 °C that exists at the bottom of Ace Lake.3,7 Several thousand years ago, Ace Lake was connected to the Southern

Introduction

640 Journal of Proteome Research 2010, 9, 640–652 Published on Web 12/01/2009

10.1021/pr900509n

 2010 American Chemical Society

research articles

Part I: The Effect of Growth Temperature 8

Ocean as a marine basin. While the lake is now a stable meromictic system with an upper-aerobic and bottom-anaerobic zone, the microbiota in the lake, including M. burtonii, are derived from the ocean. The history of the lake therefore provides a unique opportunity to study the evolution of marine microbes. Recently, the genome sequence of M. burtonii was completed, revealing a range of novel characteristics associated with its adaptation and evolution to life in a cold lake system.9 A particular feature of the genomic analysis was the important roles ascribed to genes involved in membrane and cell envelope biogenesis and genes encoding integral membrane proteins such as transporters. Despite being isolated from permanently cold waters and being able to grow at subzero temperatures, M. burtonii is capable of growth up to a maximum temperature of ∼28 °C; this temperature response classifies it as a eurypsychrophile.2,3 This capacity to tolerate a wide range of temperatures extending into the mesophilic range is typical of many microbes isolated from permanently cold environments. However, although elevated temperatures can be tolerated, temperatures that generate the fastest rates of growth (Topt) can be stressful for psychrophiles. For M. burtonii, Topt is 23 °C, a growth temperature at which the cells experience heat stress.2,3 Previous proteomic studies of M. burtonii have characterized whole-cell, soluble proteins using tandem liquid chromatography-tandem mass spectrometry (LC/LC-MS/MS) to gain a snapshot of proteins synthesized during growth at 4 °C,10 and 2D-PAGE LC-MS/MS and ICAT has been developed to initiate quantitative proteomics of cells growing at 4 vs 23 °C.11,12 In addition, LC/LC-MS/MS methods were developed for the analysis of secreted proteins.13 These studies have been restricted to growth in complex media (MFM) that provides trimethylamine (TMA) as a carbon source. In the present study, we extended proteomic analyses to include growth of M. burtonii in defined media with methanol as substrate (M-medium), enabling us to identify new proteins specific to cold adaptation under these growth conditions. By comparing proteome profiles for cells growing at 4 and 23 °C, we were able to identify core proteins involved in growth temperature specific adaptation. By comparing proteome profiles for cells growing in MFM and M-medium, we were also able to investigate responses to differences in carbon sources (TMA vs methanol) and to defined vs complex media; this work is presented in Part II14 of this study. In view of the genomic analysis highlighting the importance of cell envelope, membrane, and secreted proteins (see above), we developed new methods for analyzing less soluble proteins. To achieve quantitative proteomics with multiple growth conditions, we utilized isobaric tags for quantification (iTRAQ) enabling the nonspecific labeling of the N-terminus of expressed proteins.15 By performing quantitative proteomics on soluble, insoluble, and secreted fractions of cells grown at 4 and 23 °C in MFM and M-medium, we achieved for the first time a thorough assessment of both thermal (this work) and metabolic (Part II14) adaptation in M. burtonii.

Experimental Procedures Culture Conditions. M. burtonii (DSM 6242) cultures were grown in MFM.7,11 M-medium (pH 6.8) is based on Sowers,16 and contained: 23.38 g L-1 NaCl, 12.32 g L-1 MgSO4.6H2O, 0.76 g L-1 KCl, 0.14 g L-1 CaCl2.2H2O, 0.5 g L-1 NH4Cl, 1 mg L-1 resazurin, 10 mL vitamin solution (as for MFM), 10 mL mineral solution (as for MFM), 0.2 mL L-1 32% (v/v) HCl, 0.75 g L-1

-1

thioglycolic acid, 1.12 g L Na2HPO4.7H2O, 0.25 g L-1 -1 cysteine · HCl · H2O, and 3 g L Na2CO3. After autoclaving, Na2S (0.025% w/v final concentration) and methanol (50 mM final concentration added as two equal aliquots during growth) were added anaerobically. M. burtonii was grown in 100 mL volumes of MFM and M-medium at 4 or 23 °C under a gas phase of 80:20 N2/CO2. Cultures were inoculated 1:100 from cultures grown under the same conditions. Harvesting of M. burtonii Cultures and Fraction Separation. Three fractions were obtained from each harvested culture: supernatant (enriched for secreted proteins), insoluble (enriched for integral membrane, membrane-associated proteins, and dense macromolecular complexes), and soluble. Cultures were harvested at late logarithmic phase (optical density at 620 nm [OD620] of 0.25) by centrifugation at 3200× g for 35 min at 4 °C. The supernatant was filtered through a 15 mL Amicon centrifugal concentration unit (Millipore, Billerica, MA) with a 3 kDa cutoff by centrifugation at 5000× g, with at least two subsequent buffer exchanges with 25 mM NaHCO3. The resulting filtrate constituted the supernatant fraction.13 The insoluble and soluble fractions were separated using a method modified from Blonder et al.17 The cell pellet was suspended by vortexing in 1 mL disruption buffer (50 mM Tris.Cl; 2 mM EDTA, pH 7.2; 0.4 mM phenylmethanesulphonylfluoride). The pellet was ultrasonically disrupted on ice using a digital sonifier 250 (Branson Ultrasonics, Danbury, CT). The resulting suspension was diluted to 10 mL in ice-cold 0.1 M NaCO3 pH 11.2 and agitated on a shaker in an ice-water bath for 1 h. The suspension was transferred to 5 mL OptiSeal tubes (Beckman Coulter, Palo Alto, CA), and ultracentrifuged for 1 h at 115 000× g. The resulting supernatant was retained and filtered, as described above for the supernatant fraction, to generate the soluble fraction. The insoluble pellet was washed twice with dH2O, and then 25 mM NaHCO3 was added for a further round of ultracentrifugation (115 000× g, 20 min). The resulting supernatant was discarded. The pellet was resuspended in 2 mL of 25 mM NaHCO3 by brief sonication and concentrated in a centrifugal concentrator as described above. This homogenate was the insoluble fraction. Insoluble protein fractions required solubilization prior to trypsin digestion; this was achieved by thermally denaturing the proteins at 90 °C for 2 min, adding HPLC grade methanol to a final methanol concentration of 60% (v/v), and briefly sonicating the proteinsolvent mixture. The amount of protein in each cellular fraction was estimated at 595 nm using the method of Bradford18 with bovine serum albumin as the standard. The integrity of the cell pellet was determined by measuring the activity ratio of the cytoplasmic enzyme glutamate dehydrogenase (GDH) in the soluble and supernatant fractions.13 GDH activity was ∼1:10 between soluble and supernatant assays, for cells grown at 4 and 23 °C in MFM and M-medium, illustrating that cells under all growth conditions had remained relatively intact during processing.13 The relative abundance of the GDH protein in the soluble and supernatant fractions was also measured using iTRAQ labeling (see below) and found to be ∼1:10. Sample Preparation for iTRAQ Labeling and LC/LC-MS/ MS. The 4plex iTRAQ postincorporation labeling system (Applied Biosystems, Forster City, CA) was used according to the manufacturer’s instructions.15 Four separate comparisons were carried out for each fraction (soluble, insoluble, supernatant): 4 vs 23 °C for cells grown in MFM; 4 vs 23 °C for cells grown in M-medium; cells grown in MFM vs M-medium at 4 °C; cells Journal of Proteome Research • Vol. 9, No. 2, 2010 641

research articles grown in MFM vs M-medium at 23 °C. The 4plex iTRAQ system allows 4 samples to be labeled per LC/LC-MS/MS run. Two samples for each condition were included per run, as biological replicates. All labeling experiments were run twice (2 separate injections) using LC/LC-MS/MS, to provide a technical replicate. Six labeling experiments were performed for each culture condition, which resulted in 12 biological replicates and a total of 12 LC/LC-MS/MS runs per culture condition. Prior to iTRAQ labeling, protein samples (50 µg) were reduced, alkylated, and digested. Samples of soluble, supernatant, and insoluble fractions were treated with 1 µL of 1% sodium dodecyl sulfate (SDS), then reduced with either 2 µL of tris(2-carboxyethyl)phosphine (1 mM soluble and supernatant fractions) or 5 µL of 0.1 M dithiothreitol (insoluble fraction) and incubated in the dark at 60 °C for 60 min. All samples were alkylated with iodoacetamide (2 mM for soluble and supernatant fractions, 15 mM for insoluble fraction), and incubated in the dark at 60 °C for 60 min. Supernatant and soluble protein fractions were then digested using 0.5 µg trypsin per 50 µg protein (sequencing grade porcine modified trypsin; Promega, Madison, WI) and incubated for 16 h at 37 °C. Insoluble protein fractions required solubilization prior to trypsin digestion; this was achieved by thermally denaturing the proteins at 90 °C for 2 min, adding HPLC grade methanol to a final methanol concentration of 60% (v/v), and briefly sonicating the proteinsolvent mixture. Trypsin was added to the mixture (2.5 µg trypsin per 50 µg protein), which was incubated for 5 h at 37 °C. The samples were then labeled with the iTRAQ labeling system according to the manufacturer’s instructions, and then combined. To remove any unbound iTRAQ labels, trypsin, SDS, solvents, etc., the combined samples were cleaned-up using strong cation exchange (SCX) chromatography followed by reverse phase (RP) chromatography. Soluble and supernatant fraction samples could proceed directly to SCX, but insoluble fraction samples were vacuum-dried, and sample pellets were resuspended in 5 mL of load buffer (10 mM KH2PO4, 25% acetonitrile [ACN], pH 3.0) and centrifuged at 5000× g for 5 min at 4 °C and the supernatant was retained for SCX. An Applied Biosystems Opti-Lynx cartridge/holder and a syringe pump (KD Scientific, Holliston, MA) at a flow rate of 9.5 mL h-1 were used for SCX. The eluted peptide solutions were vacuum-dried, and the pellet resuspended in 500 µL of 0.2% heptafluorobutyric acid (HFBA). A RP peptide macrotrap (Microm Bioresources, Auburn, CA) was used and prepared by washing with 1 mL ACN, then 1 mL ACN:0.25% formic acid (1:1), followed by equilibration using 1 mL of 0.2% HFBA. The peptides were loaded, and the column washed with 1.5 mL 0.2% HFBA. Peptides were eluted with 500 µL ACN:0.25% (v/v) formic acid (50:50), followed by 500 µL ACN. For peptides derived from all three cellular fractions, the eluent was vacuumdried and the pellet dissolved in 50 µL 1% formic acid, 0.05% HFBA. Mass Spectrometry and Data Analysis. Solubilized peptides were separated online with automated high throughput SCX and nano C18 LC using an Ultimate HPLC, Switchos and Famos autosampler system (LC-Packings, Amsterdam, Netherlands), and eluted stepwise with increasing concentrations of ammonium acetate (11). Samples were analyzed using an API QStar Pulsar i hybrid tandem mass spectrometer (Applied Biosystems, Foster City, CA) operated in information dependent acquisition mode. Peptides were eluted using a linear acetonitrile gradient. A time-of-flight (TOF) MS survey scan was 642

Journal of Proteome Research • Vol. 9, No. 2, 2010

Williams et al. acquired (m/z 350-1700, 1s). The 3 largest multiply charged ions (counts >25) were sequentially selected by the quadrupole for MS-MS analysis.19 Tandem mass spectra were accumulated for 2.5s (m/z 65-2000). Peak lists were generated using Mascot Distiller (Matrix Science, London, UK) using the default parameters. MS data were searched using Mascot against an in silico digestion of the local M. burtonii protein FASTA database created in Mascot, with appropriate parameters (tryptic digestion; variable modifications of carbamidomethylation; methionine oxidation; 4-plex iTRAQ; peptide ion mass tolerance (0.25 Da; fragment ion mass tolerance (0.2 Da; maximum number of missed cleavages set to 1). A typical fragmentation spectrum is shown in Figure S1 (Supporting Information). A decoy database created by randomizing the M. burtonii local FASTA database, created with Mascot, was also searched with the same parameters. All spectra which matched the databases with a Mowse score of 30 were manually inspected to ensure ion progressions of 4 or more consecutive ions of a single class (e.g., y- or b-type ions); any matches that did not meet these criteria were rejected. From the identifications in both the M. burtonii database and the decoy database it was possible to calculate the false discovery rate (FDR) for any given experimental run.20 Any experiment that had a FDR > 0.02 was rejected. Identifications were then moved to a database for further processing. iTRAQ abundance data were determined using ProQuant software (Build 7051; Applied Biosystems, Foster City, CA) and searched against the local M. burtonii interrogator database (MS tolerance 0.2 Da; MS/MS tolerance 0.15 Da; store hits with confidence >99%; charge range for intermediate precursors of g2, e4; maximum number of missed cleavages 1; modifications: iTRAQ reagents, acetamidomethyl-cysteine, methionine sulfoxide, with maximum number of modifications set to 8; correction factors entered from the iTRAQ certificate of analysis). Data were visualized using ProGroup Viewer version 1.0.5 (Applied Biosystems, Foster City, CA). We used SPSS release 17.0.1 (IBM, Armonk, NY) to compare data variability in biological vs technical replicates of proteins identified by iTRAQ labeling and ProQuant. These demonstrated that data were reproducible across biological and technical replicates; representative scatter plots are shown in Figure S2 (Supporting Information). Data were inspected for variation in abundance (p < 0.05), with error factor (EF) values 1.5 fold) in the expressed proteome. Figure S1 shows a typical fragmentation spectrum for a peptide derived from Mbur_1950 (DEAD box RNA helicase). Figure S2 shows a scatter plot to demonstrate reproducibility across samples (biological replicates) and injections (technical replicates). This material is available free of charge via the Internet at http://pubs.acs.org. References (1) Murray, A. E.; Grzymski, J. J. Diversity and genomics of Antarctic marine micro-organisms. Philos. Trans. R. Soc. Lond. B: Biol. Sci. 2007, 362, 2259–2271. (2) Feller, G.; Gerday, C. Psychrophilic enzymes: hot topics in cold adaptation. Nat. Rev. Microbiol. 2003, 1, 200–208. (3) Cavicchioli, R. Cold-adapted archaea. Nat. Rev. Microbiol. 2006, 4, 331–343. (4) Cavicchioli, R.; Thomas, T.; Curmi, P. M. G. Cold stress response in Archaea. Extremophiles 2000, 4, 321–331. (5) Takai, K.; Nakamura, K.; Toki, T.; Tsunogai, U.; Miyazaki, M.; Miyazaki, J.; Hirayama, H.; Nakagawa, S.; Nunoura, T.; Horikoshi, K. Cell proliferation at 122°C and isotopically heavy CH4 production by a hyperthermophilic methanogen under high-pressure cultivation. Proc. Natl. Acad. Sci. U.S.A. 2008, 105, 10949–10954. (6) Reid, I. N.; Sparks, W. B.; Lubnow, S.; McGrath, M.; Livio, M.; Valenti, J.; Sowers, K. A.; Shukla, H. D.; MacAuley, S.; Miller, T.; Suvanasuthi, R.; Belas, R.; Colman, A.; Robb, F. T.; DasSarma, P.; Mu ¨ ller, J. A.; Coker, J. A.; Cavicchioli, R.; Chen, F.; DasSarma, S. (2006) Terrestrial models for extraterrestrial life: methanogens and halophiles at Martian temperatures. Int. J. Astrobiol. 2006, 5, 89– 97. (7) Franzmann, P. D.; Springer, N.; Ludwig, W.; Conway De Macario, E.; Rohde, M. A methanogenic archaeon from Ace Lake, Antarctica: Methanococcoides burtonii sp. nov. System. Appl. Microbiol. 1992, 15, 573–581. (8) Rankin, L. M.; Gibson, J. A. E.; Franzmann, P. D.; Burton, H. R. The chemical stratification and microbial communities of Ace Lake, Antarctica: a review of the characteristics of a marine-derived meromictic lake. Polarforschung 1999, 66, 33–52. (9) Allen, M. A.; Lauro, F. M.; Williams, T. J.; Burg, D,; Siddiqui, K. S.; De Franciscii, D.; Chong, K. W. Y.; Pilak, O.; Chew, H. H.; De Maere, M. Z.; Ting, L.; Katrib, M.; Ng, C.; Sowers, K. R.; Galeprin, M. Y.; Anderson, I. J.; Ivanova, N.; Dalin, E.; Martinez, M.; Lapidus, A.; Hauser, L.; Land, M.; Thomas, T.; Cavicchioli, R. The genome sequence of the psychrophilic archaeon, Methanococcoides burtonii: the role of genome evolution in cold-adaptation. ISME J. 2009, 3, 1012–1035. (10) Goodchild, A.; Raftery, M.; Saunders, N. F. W.; Guilhaus, M.; Cavicchioli, R. Biology of the cold adapted archaeon, Methano-

650

Journal of Proteome Research • Vol. 9, No. 2, 2010

Williams et al.

(11)

(12)

(13)

(14) (15)

(16)

(17)

(18) (19)

(20) (21)

(22)

(23)

(24)

(25)

(26)

(27) (28)

(29) (30)

coccoides burtonii determined by proteomics using liquid chromatography-tandem mass spectrometry. J. Proteome Res. 2004, 3, 1164–1176. Goodchild, A.; Saunders, N. F. W.; Ertan, H.; Raftery, M.; Guilhaus, M.; Curmi, P. M. G.; Cavicchioli, R. A proteomic determination of cold adaptation in the Antarctic archaeon, Methanococcoides burtonii. Mol. Microbiol. 2004, 53, 309–321. Goodchild, A.; Raftery, M.; Saunders, N. F. W.; Guilhaus, M.; Cavicchioli, R. Cold adaptation of the Antarctic archaeon, Methanococcoides burtonii assessed by proteomics using ICAT. J. Proteome. Res. 2005, 4, 473–480. Saunders, N. F. W.; Ng, C.; Raftery, M.; Guilhaus, M.; Goodchild, A.; Cavicchioli, R. Proteomic and computational analysis of secreted proteins with type I signal peptides from the Antarctic archaeon Methanococcoides burtonii. J. Proteome Res. 2006, 5, 2457–2464. Williams, T. J.; Burg, D. W.; Ertan, H.; Raftery, M. J.; Poljak, A.; Guilhaus, M.; Cavicchioli, R. J. Proteome Res. 2010, 9, DOI: 10.1021/ pr9005102. Ross, P. L.; Huang, Y. N.; Marchese, J. N.; Williamson, B.; Parker, K.; Hattan, S.; Khainovski, N.; Pillai, S.; Dey, S.; Daniels, S.; Purkayastha, S.; Juhasz, P.; Martin, S.; Bartlet-Jones, M.; He, F.; Jacobson, A.; Pappin, D. J. Multiplexed protein quantitation in Saccharomyces cerevisiae using amine-reactive isobaric tagging reagents. Mol. Cell. Proteomics 2004, 3, 1154–1169. Sowers, K. R. Methanogenic Archaea: an overview. In: Archaea. A Laboratory Manual; Robb, F. T., Place, A. R., Sowers, K. R., Schreier, H. J., DasSarma, S., Fleischmann, E. M., Eds.; Cold Spring Harbor Laboratory Press: Plainview, 1995; pp 3-13. Blonder, J.; Goshe, M. B.; Moore, R. J.; Pasa-Tolic, L.; Masselon, C. D.; Lipton, M. S.; Smith, R. D. Enrichment of integral membrane proteins for proteomic analysis using liquid chromatographytandem mass spectrometry. J. Proteome Res. 2002, 1, 351–360. Bradford, M. M. (1976). A rapid and sensitive method for quantitation of microgram quantities of protein utilizing the principle of protein-dye-binding. Anal. Biochem. 1976, 72, 248–254. Leuko, S.; Raftery, M.; Burns, B.; Walter, M.; Neilan, B. Global protein-level responses of Halobacterium salinarum NRC-1 to prolonged changes in external sodium chloride concentrations. J. Proteome Res. 2009, 8, 2218–2225. Kall, L.; Storey, J. D.; MacCoss, M. J.; Noble, W. S. Assigning significance to peptides identified by tandem mass spectrometry using decoy databases. J. Proteome Res. 2008, 7, 29–34. Zhou, L.; Beuerman, R. W.; Huang, L.; Barathi, A.; Foo, Y. H.; Li, S. F. Y.; Chew, F. T.; Tan, D. Proteomic analysis of rabbit tear fluid: Defensin levels after an experimental corneal wound are correlated to wound closure. Proteomics 2007, 7, 3194–3206. Saunders, N. F. W.; Goodchild, A.; Raftery, M.; Guilhaus, M.; Curmi, P. M.; Cavicchioli, R. Predicted roles for Hypothetical proteins in the low-temperature expressed proteome of the Antarctic archaeon Methanococcoides burtonii. J. Proteome Res. 2005, 4, 464–472. Jing, H.; Takagi, J.; Liu, J. H.; Lindgren, S.; Zhang, R. G.; Joachimiak, A.; Wang, J. H.; Springer, T. A. Archaeal surface layer proteins contain beta propeller, PKD, and beta helix domains and are related to metazoan cell surface proteins. Structure 2002, 10, 1453– 1464. Ko¨nig, H.; Rachel, R.; Claus, H. Proteinaceous surface layers of Archaea: ultrastructure and biochemistry. In: Cavicchioli, R. (Ed.) Archaea. Molecular and Cellular Biology; ASM Press: Washington, DC, 2007; pp 315-340. Yang, L.; Zhou, D.; Liu, X.; Han, H.; Zhan, L.; Guo, Z.; Zhang, L.; Qin, C.; Wong, H. C.; Yang, R. Cold-induced gene expression profiles of Vibrio parahaemolyticus: a time-course analysis. FEMS Microbiol. Lett. 2009, 291, 50–58. Weinberg, M. V.; Schut, G. J.; Brehm, S.; Datta, S.; Adams, M. W. Cold shock of a hyperthermophilic archaeon: Pyrococcus furiosus exhibits multiple responses to a suboptimal growth temperature with a key role for membrane-bound glycoproteins. J. Bacteriol. 2005, 187, 336–348. Takeichi, M. Cadherins: a molecular family important in selective cell-cell adhesion. Annu. Rev. Biochem. 1990, 59, 237–252. Kelly, G.; Prasannan, S.; Daniell, S.; Fleming, K.; Frankel, G.; Dougan, G.; Connerton, I.; Matthews, S. Structure of the celladhesion fragment of intimin from enteropathogenic Escherichia coli. Nat. Struct. Biol. 1999, 6, 313–318. Peer, A.; Smith, S. P.; Bayer, E. A.; Lamed, R.; Borovok, I. Noncellulosomal cohesin- and dockerin-like modules in the three domains of life. FEMS Microbiol. Lett. 2008, 291, 1–16. Shoham, Y.; Lamed, R.; Bayer, E. A. The cellulosome concept as an efficient microbial strategy for the degradation of insoluble polysaccharides. Trends. Microbiol. 1999, 7, 275–281.

Part I: The Effect of Growth Temperature (31) Galagan, J. E.; Nusbaum, C.; Roy, A.; Endrizzi, M. G.; Macdonald, P.; FitzHugh, W.; Calvo, S.; Engels, R.; Smirnov, S.; Atnoor, D.; Brown, A.; Allen, N.; Naylor, J.; Stange-Thomann, N.; DeArellano, K.; Johnson, R.; Linton, L.; McEwan, P.; McKernan, K.; Talamas, J.; Tirrell, A.; Ye, W.; Zimmer, A.; Barber, R. D.; Cann, I.; Graham, D. E.; Grahame, D. A.; Guss, A. M.; Hedderich, R.; Ingram-Smith, C.; Kuettner, H. C.; Krzycki, J. A.; Leigh, J. A.; Li, W.; Liu, J.; Mukhopadhyay, B.; Reeve, J. N.; Smith, K.; Springer, T. A.; Umayam, L. A.; White, O.; White, R. H.; de Macario, E. C.; Ferry, J. G.; Jarrell, K. F.; Jing, H.; Macario, A. J.; Paulsen, I.; Pritchett, M.; Sowers, K. R.; Swanson, R. V.; Zinder, S. H.; Lander, E.; Metcalf, W. W.; Birren, B. The genome of M. acetivorans reveals extensive metabolic and physiological diversity. Genome Res. 2002, 12, 532– 542. (32) Nichols, D. S.; Miller, M. R.; Davies, N. W.; Goodchild, A.; Raftery, M.; Cavicchioli, R. Cold adaptation in the Antarctic archaeon Methanococcoides burtonii involves membrane lipid unsaturation. J. Bacteriol. 2004, 186, 8508–8515. (33) Neuwald, A. F.; Aravind, L.; Spouge, J. L.; Koonin, E. V. AAA+: A class of chaperone-like ATPases associated with the assembly, operation, and disassembly of protein complexes. Genome Res. 1999, 9, 27–43. (34) Sowers, K. R.; Gunsalus, R. P. Halotolerance in Methanosarcina spp.: Role of Nε-acetyl-β-lysine, R-glutamate, glycine betaine and K+ as compatible solutes for osmotic adaptation. Appl. Environ. Microbiol. 1995, 61, 4382–4388. (35) Thomas, T.; Kumar, N.; Cavicchioli, R. Effects of ribosomes and intracellular solutes on activities and stabilities of elongation factor 2 proteins from psychrotolerant and thermophilic methanogens. J. Bacteriol. 2001, 183, 1974–1982. (36) Stricker, J.; Maddox, P.; Salmon, E. D.; Erickson, H. P. Rapid assembly dynamics of the Escherichia coli FtsZ-ring demonstrated by fluorescence recovery after photobleaching. Proc. Natl. Acad. Sci. U.S.A. 2002, 99, 3171–3175. (37) Samson, R. Y.; Reeve, J. N. DNA binding proteins and chromatin. In Archaea. Molecular and Cellular Biology; Cavicchioli, R., Ed.; ASM Press: Washington, DC, 2007; pp 110-119. (38) Chartier, F.; Laine, B.; Sautiere, P. Characterization of the chromosomal protein MC1 from the thermophilic archaebacterium Methanosarcina sp. CHTI 55 and its effect on the thermal stability of DNA. Biochim. Biophys. Acta 1988, 951, 149–156. (1988). (39) Thompson, S. R.; Wadhams, G. H.; Armitage, J. P. The positioning of cytoplasmic protein clusters in bacteria. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 8209–8214. (40) Saunders, N. F. W.; Thomas, T.; Curmi, P. M.; Mattick, J. S.; Kuczek, E.; Slade, R.; Davis, J.; Franzmann, P. D.; Boone, D.; Rusterholtz, K.; Feldman, R.; Gates, C.; Bench, S.; Sowers, K.; Kadner, K.; Aerts, A.; Dehal, P.; Detter, C.; Glavina, T.; Lucas, S.; Richardson, P.; Larimer, F.; Hauser, L.; Land, M.; Cavicchioli, R. Mechanisms of thermal adaptation revealed from the genomes of the Antarctic archaea Methanogenium frigidum and Methanococcoides burtonii. Genome Res. 2003, 13, 1580–1588. (41) Lim, J.; Thomas, T.; Cavicchioli, R. Low temperature regulated DEAD-box RNA helicase from the Antarctic archaeon, Methanococcoides burtonii. J. Mol. Biol. 2000, 297, 553–567. (42) Jones, P. G.; Mitta, M.; Kim, Y.; Jiang, W.; Inouye, M. Cold-shock induces a major ribosomal associated protein that unwinds double stranded RNA in Escherichia coli. Proc. Natl. Acad. Sci. U.S.A. 1996, 93, 76–80. (43) Prud’homme-Ge´ne´reux, A.; Beran, R. K.; Iost, I.; Ramey, C. S.; Mackie, G. A.; Simons, R. W. Physical and functional interactions among RNase E, polynucleotide phosphorylase and the cold-shock protein, CsdA: evidence for a ‘cold shock degradosome’. Mol. Microbiol. 2004, 54, 1409–1421. (44) El-Fahmawi, B.; Owttrim, G. W. Polar-biased localization of the cold stress-induced RNA helicase, CrhC, in the cyanobacterium Anabaena sp. strain PCC 7120. Mol. Microbiol. 2003, 50, 1439– 1448. (45) Giaquinto, L.; Curmi, P. M.; Siddiqui, K. S.; Poljak, A.; DeLong, E.; DasSarma, S.; Cavicchioli, R. (2007) Structure and function of cold shock proteins in archaea. J. Bacteriol. 2007, 189, 5738–5748. (46) Anantharaman, V.; Koonin, E. V.; Aravind, L. TRAM, a predicted RNA-binding domain, common to tRNA uracil methylation and adenine thiolation enzymes. FEMS Microbiol. Lett. 2001, 197, 215– 221. (47) Jones, P. G.; Inouye, M. The cold-shock response - a hot topic. Mol. Microbiol. 1994, 11, 811–818. (48) Yamanaka, K.; Fang, L.; Inouye, M. The CspA family in Escherichia coli: multiple gene duplication for stress adaptation. Mol. Microbiol. 1998, 27, 247–255.

research articles (49) Christiansen, J. K.; Larsen, M. H.; Ingmer, H.; Søgaard-Andersen, L.; Kallipolitis, B. H. The RNA-binding protein Hfq of Listeria monocytogenes: role in stress tolerance and virulence. J. Bacteriol. 2004, 186, 3355–3362. (50) Even, S.; Pellegrini, O.; Zig, L.; Labas, V.; Vinh, J.; Bre´chemmierBaey, D.; Putzer, H. Ribonucleases J1 and J2: two novel endoribonucleases in B. subtilis with functional homology to E. coli RNase E. Nucleic Acids Res. 2005, 33, 2141–2152. (51) Haldenby, S.; White, M. F.; Allers, T. RecA family proteins in archaea: RadA and its cousins. Biochem. Soc. Trans. 2009, 37, 102– 107. (52) Moldovan, G. L.; Pfander, B.; Jentsch, S. PCNA, the maestro of the replication fork. Cell 2007, 129, 665–679. (53) Fan, L.; Arvai, A. S.; Cooper, P. K.; Iwai, S.; Hanaoka, F.; Tainer, J. A. Conserved XPB core structure and motifs for DNA unwinding: implications for pathway selection of transcription or excision repair. Mol. Cell 2006, 22, 27–37. (54) Buckel, W.; Golding, B. T. Radical enzymes in anaerobes. Annu. Rev. Microbiol. 2006, 60, 27–49. (55) Scho¨neich, C.; Asmus, K. D. Reaction of thiyl radicals with alcohols, ethers and polyunsaturated fatty acids: a possible role of thiyl free radicals in thiol mutagenesis. Radiat. Environ. Biophys. 1990, 29, 263–271. (56) Scho¨neich, C. Mechanisms of protein damage induced by cysteine thiyl radical formation. Chem. Res. Toxicol. 2008, 21, 1175–1179. (57) Ferreri, C.; Kratzsch, S.; Landi, L.; Brede, O. Thiyl radicals in biosystems: effects on lipid structures and metabolisms. Cell. Mol. Life Sci. 2006, 62, 834–847. (58) Wang, S. C.; Frey, P. A. S-adenosylmethionine as an oxidant: the radical SAM superfamily. Trends Biochem. Sci. 2007, 32, 101–110. (59) Ugulava, N. B.; Frederick, K. K.; Jarrett, J. T. Control of adenosylmethionine-dependent radical generation in biotin synthase: A kinetic and thermodynamic analysis of substrate binding to active and inactive forms of BioB. Biochemistry 2003, 42, 2708–2719. (60) Theil, E. C. Coordinating responses to iron and oxygen stress with DNA and mRNA promoters: the ferritin story. Biometals 2007, 20, 513–521. (61) Nachin, L.; Nannmark, U.; Nystro¨m, T. Differential roles of the universal stress proteins of Escherichia coli in oxidative stress resistance, adhesion, and motility. J. Bacteriol. 2005, 187, 6265– 6272. (62) Kvint, K.; Nachin, L.; Diez, A.; Nystrom, T. The bacterial universal stress protein: function and regulation. Curr. Opin. Microbiol. 2003, 6, 140–145. (63) Medina-Silva, R.; Barros, M. P.; Galhardo, R. S.; Netto, L. E.; Colepicolo, P.; Menck, C. F. Heat stress promotes mitochondrial instability and oxidative responses in yeast deficient in thiazole biosynthesis. Res. Microbiol. 2006, 157, 275–281. (64) Li, Q.; Li, L.; Rejtar, T.; Karger, B. L.; Ferry, J. G. Proteome of Methanosarcina acetivorans Part I: an expanded view of the biology of the cell. J. Proteome Res. 2005, 4, 112–128. (65) Londei, P. Translation. In Archaea. Molecular and Cellular Biology; Cavicchioli, R., Ed.; ASM Press: Washington, DC, 2007; pp 175208. (66) Ring, G.; Eichler, J. Membrane binding of ribosomes occurs at SecYE-based sites in the Archaea Haloferax volcanii. J. Mol. Biol. 2004, 336, 997–1010. (67) Francoleon, D. R.; Boontheung, P.; Yang, Y.; Kim, U.; Ytterberg, A. J.; Denny, P. A.; Denny, P. C.; Loo, J. A.; Gunsalus, R. P.; Ogorzalek, Loo; R. R. S-layer, surface-accessible, and concanavalin A binding proteins of Methanosarcina acetivorans and Methanosarcina mazei. J. Proteome Res. 2009, 8, 1972–1982. (68) Caldas, T. D.; El Yaagoubi, A.; Richarme, G. Chaperone properties of bacterial elongation factor EF-Tu. J. Biol. Chem. 1998, 273, 11478–11482. (69) Caldas, T.; Laalami, S.; Richarme, G. J. Chaperone properties of bacterial elongation factor EF-G and initiation factor IF2. Biol. Chem. 2000, 275, 855–860. (70) Kovacs, D.; Rakacs, M.; Agoston, B.; Lenkey, K.; Semrad, K.; Schroeder, R.; Tompa, P. Janus chaperones: assistance of both RNA- and protein-folding by ribosomal proteins. FEBS Lett. 2009, 583, 88–92. (71) Kandror, O.; Goldberg, A. L. Trigger factor is induced upon cold shock and enhances viability of Escherichia coli at low temperatures. Proc. Natl. Acad. Sci. U.S.A. 1997, 94, 4978–4981. (72) Ideno, A.; Yoshida, T.; Iida, T.; Furutani, M.; Maruyama, T. FK506binding protein of the hyperthermophilic archaeum, Thermococcus sp. KS-1, a cold-shock inducible peptidyl-prolyl cis-trans isomerase with activities to trap and refold denatured proteins. Biochem. J. 2001, 357, 465–471.

Journal of Proteome Research • Vol. 9, No. 2, 2010 651

research articles (73) Phadtare, S. Recent developments in bacterial cold-shock response. Curr. Issues Mol. Biol. 2004, 6, 125–136. (74) Maupin-Furlow, J. A.; Humbard, M. A.; Kirkland, P. A.; Li, W.; Reuter, C. J.; Wright, A. J.; Zhou, G. (2006). Proteasomes from structure to function: perspectives from Archaea. Curr. Top. Dev. Biol. 2006, 75, 125–169. (75) Roberts, T. H.; Hejgaard, J.; Saunders, N. F. W.; Cavicchioli, R.; Curmi, P. M. Serpins in unicellular Eukarya, Archaea, and Bacteria: sequence analysis and evolution. J. Mol. Evol. 2004, 59, 437–447.

652

Journal of Proteome Research • Vol. 9, No. 2, 2010

Williams et al. (76) Tavernarakis, N.; Driscoll, M.; Kyrpides, N. C. The SPFH domain: implicated in regulating targeted protein turnover in stomatins and other membrane-associated proteins. Trends Biochem. Sci. 1999, 24, 425–427. (77) Sakoh, M.; Ito, K.; Akiyama, Y. Proteolytic activity of HtpX, a membrane-bound and stress-controlled protease from Escherichia coli. J. Biol. Chem. 2005, 280, 33305–33310.

PR900509N