Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 20 de 112
Filtrar
Más filtros

Banco de datos
País/Región como asunto
Tipo del documento
Intervalo de año de publicación
1.
EMBO Rep ; 25(2): 770-795, 2024 Feb.
Artículo en Inglés | MEDLINE | ID: mdl-38182816

RESUMEN

DExD/H-box helicases are crucial regulators of RNA metabolism and antiviral innate immune responses; however, their role in bacteria-induced inflammation remains unclear. Here, we report that DDX5 interacts with METTL3 and METTL14 to form an m6A writing complex, which adds N6-methyladenosine to transcripts of toll-like receptor (TLR) 2 and TLR4, promoting their decay via YTHDF2-mediated RNA degradation, resulting in reduced expression of TLR2/4. Upon bacterial infection, DDX5 is recruited to Hrd1 at the endoplasmic reticulum in an MyD88-dependent manner and is degraded by the ubiquitin-proteasome pathway. This process disrupts the DDX5 m6A writing complex and halts m6A modification as well as degradation of TLR2/4 mRNAs, thereby promoting the expression of TLR2 and TLR4 and downstream NF-κB activation. The role of DDX5 in regulating inflammation is also validated in vivo, as DDX5- and METTL3-KO mice exhibit enhanced expression of inflammatory cytokines. Our findings show that DDX5 acts as a molecular switch to regulate inflammation during bacterial infection and shed light on mechanisms of quiescent inflammation during homeostasis.


Asunto(s)
Adenina , Infecciones Bacterianas , Receptor Toll-Like 2 , Animales , Ratones , Adenina/análogos & derivados , Inflamación/genética , Metiltransferasas/genética , Receptor Toll-Like 2/genética , Receptor Toll-Like 4/genética
2.
Small ; 19(9): e2205898, 2023 Mar.
Artículo en Inglés | MEDLINE | ID: mdl-36534903

RESUMEN

The majority of commercial polyolefins are produced by coordination polymerization using early or late transition metal catalysts. Molecular catalysts containing these transition metals (Ti, Zr, Cr, Ni, and Fe, etc.) are loaded on supports for controlled polymerization behavior and polymer morphology in slurry or gas phase processes. Within the last few years, metal-organic frameworks (MOFs), a class of unique porous crystalline materials constructed from metal ions/clusters and organic ligands, have been designed and utilized as excellent supports for heterogeneous polymerization catalysis whose high density and uniform distribution of active sites would benefit the modulations of molecular weight distributions of high-performance olefin oligomers and (co)polymers. Impressive efforts have been made to modulate the microenvironment surrounding the active centers at the atomic level for improved activities of MOFs-based catalysts and controlled selectivity of olefin insertion. This review aims to draw a comprehensive picture of MOFs for coordination olefin oligomerization and (co)polymerization in the past decades with respect to different transition metal active centers, various incorporation sites, and finally microenvironment modulation. In consideration of more efforts are needed to overcome challenges for further industrial and commercial application, a brief outlook is provided.

3.
Annu Rev Microbiol ; 71: 459-477, 2017 09 08.
Artículo en Inglés | MEDLINE | ID: mdl-28697670

RESUMEN

Dormant Bacillales and Clostridiales spores begin to grow when small molecules (germinants) trigger germination, potentially leading to food spoilage or disease. Germination-specific proteins sense germinants, transport small molecules, and hydrolyze specific bonds in cortex peptidoglycan and specific proteins. Major events in germination include (a) germinant sensing; (b) commitment to germinate; (c) release of spores' depot of dipicolinic acid (DPA); (d) hydrolysis of spores' peptidoglycan cortex; and (e) spore core swelling and water uptake, cell wall peptidoglycan remodeling, and restoration of core protein and inner spore membrane lipid mobility. Germination is similar between Bacillales and Clostridiales, but some species differ in how germinants are sensed and how cortex hydrolysis and DPA release are triggered. Despite detailed knowledge of the proteins and signal transduction pathways involved in germination, precisely what some germination proteins do and how they do it remain unclear.


Asunto(s)
Bacillales/crecimiento & desarrollo , Clostridiales/crecimiento & desarrollo , Peptidoglicano/metabolismo , Esporas/crecimiento & desarrollo , Pared Celular/metabolismo , Fluidez de la Membrana , Lípidos de la Membrana/metabolismo , Ácidos Picolínicos/metabolismo , Agua/metabolismo
4.
Anal Chem ; 93(3): 1443-1450, 2021 01 26.
Artículo en Inglés | MEDLINE | ID: mdl-33369381

RESUMEN

A prophage comprises a bacteriophage genome that has integrated into a host bacterium's DNA, which generally permits the cell to grow and divide normally. However, the prophage can be induced by various stresses, or induction can occur spontaneously. After prophage induction, viral replication and production of endolysins begin until the cell lyses and phage particles are released. However, the heterogeneity of prophage induction and lysis of individual cells in a population and the dynamics of a cell undergoing lysis by prophage induction have not been fully characterized. Here, we used Raman tweezers and live-cell phase-contrast microscopy to characterize the Raman spectral and cell length changes that occur during the lysis of individual Bacillus subtilis cells from spores that carry PBSX prophage during spores' germination, outgrowth, and then vegetative growth. Major findings of this work are as follows: (i) After addition of xylose to trigger prophage induction, the intensities of Raman spectral bands associated with nucleic acids of single cells in induced cultures gradually fell to zero, in contrast to the much smaller changes in protein band intensities and no changes in nucleic acid bands in uninduced cultures; (ii) the nucleic acid band intensities from an individual induced cell exhibited a rapid decrease, following a long lag period; (iii) after the addition of nutrient-rich medium with xylose, single spores underwent a long period (228 ± 41.4 min) for germination, outgrowth, and vegetative growth, followed by a short period of cell burst in 1.5 ± 0.8 min at a cell length of 8.2 ± 5.5 µm; (iv) the latent time (Tlatent) between the addition of xylose and the start of cell burst was heterogeneous in cell populations; however, the period (ΔTburst) from the latent time to the completion of cell lysis was quite small; (v) in a poor medium with l-alanine alone, addition of xylose caused prophage induction following spore germination but with longer Tlatent and ΔTburst times and without cell elongation; (vi) spontaneous prophage induction and lysis of individual cells from spores in a minimal nutrient medium were observed without xylose addition, and cell length prior to cell lysis was ∼4.1 µm, but spontaneous prophage induction was not observed in a rich medium; (vii) in a rich medium, addition of xylose at a time well after spore germination and outgrowth significantly shortened the average Tlatent time. The results of this study provide new insights into the heterogeneity and dynamics of lysis of individual B. subtilis cells derived from spores upon prophage induction.


Asunto(s)
Bacillus subtilis/citología , Análisis de la Célula Individual , Esporas Bacterianas/crecimiento & desarrollo , Bacillus subtilis/metabolismo , Microscopía de Contraste de Fase , Pinzas Ópticas , Espectrometría Raman , Esporas Bacterianas/química , Esporas Bacterianas/metabolismo
5.
Anal Chem ; 92(1): 1326-1332, 2020 01 07.
Artículo en Inglés | MEDLINE | ID: mdl-31793766

RESUMEN

Confocal Raman microscopy is a powerful method for nondestructive and noninvasive detection of chemicals with high spatial resolution, but its long acquisition time hinders its applications in large-scale monitoring of fast dynamics. Here, we report the development of a compressive sensing technique for single-acquisition multifocal Raman spectroscopy, which is capable of improving the speed of conventional confocal Raman spectroscopy by 2-3 orders of magnitude. A sample is excited with a 2-D multifocus pattern, and the Raman scatterings from the multiple foci were projected onto the spectrometer's entrance in a 2-D array. The superimposed spectra within each row of the array were processed with an algorithm such that the spectra from the individual foci were retrieved in a single acquisition and with reduced noise. The performances of the developed technique were demonstrated by parallel Raman spectroscopy of multiple individual particles as well as by single-acquisition confocal Raman imaging of a large scale with high spatial resolution when combined with spatially sparse sampling. The technique is expected to find wide applications in investigating fast dynamics in large-scale biological systems.

6.
Yi Chuan ; 42(6): 536-547, 2020 Jun 20.
Artículo en Zh | MEDLINE | ID: mdl-32694112

RESUMEN

Cardiac hypertrophy is a compensatory response that occurs as a result of increased hemodynamic requirement in peripheral tissues. In the process of cardiac hypertrophy, the expression of different types of genes in different stages is transcriptionally regulated by multiple-level physiological and pathological signals. Histone acetylation, as the most extensive post-translational modification, is closely controlled by the antagonistic histone acetyltransferases (HAT) and histone deacetylases (HDACs). Recent studies have shown that HDACs, as a family of enzymes that inhibit transcription and contain highly conserved deacetylase domains, regulate gene expression during cardiac hypertrophy through a variety of pathways. In this review, we mainly summarize the research progress on histone deacetylase in cardiac hypertrophy. By elucidating the role and molecular mechanism of different HDACs in cardiac hypertrophy, it provides new ideas for the treatment of different types of cardiac hypertrophy and heart failure, and molecular targets for new drug design.


Asunto(s)
Cardiomegalia , Histona Desacetilasas , Acetilación , Cardiomegalia/genética , Histona Acetiltransferasas , Inhibidores de Histona Desacetilasas , Histona Desacetilasas/genética , Histona Desacetilasas/metabolismo , Humanos , Procesamiento Proteico-Postraduccional
7.
Zhongguo Zhong Yao Za Zhi ; 45(23): 5577-5588, 2020 Dec.
Artículo en Zh | MEDLINE | ID: mdl-33496095

RESUMEN

Unraveling the genetic basis of medicinal plant metabolism and developmental traits is a long-standing goal for pharmacologists and plant biologists. This paper discusses the definition of molecular genetics of medicinal plants, which is an integrative discipline with medicinal plants as the research object. This discipline focuses on the heredity and variation of medicinal plants, and elucidates the relationship between the key traits of medicinal plants(active compounds, yield, resistance, etc.) and genotype, studies the structure and function, heredity and variation of medicinal plant genes mainly at molecular level, so as to reveal the molecular mechanisms of transmission, expression and regulation of genetic information of medicinal plants. Specifically, we emphasize on three major aspects of this discipline.(1)Individual and population genetics of medicinal plants, this part mainly highlights the genetic mechanism of the domestication, the individual genomics at the species level, and the formation of genetic diversity of medicinal plants.(2)Elucidation of biosynthetic pathways of active compounds and their evolutionary significance. This part summarizes the biosynthesis, diversity and molecular evolution of active compounds in medicinal plants.(3) Molecular mechanisms that shaping the key agronomic traits by internal and external factors. This part focuses on the accumulation and distribution of active compounds within plants and the regulation of metabolic network by environmental factors. Finally, we prospect the future direction of molecular genetics of medicinal plants based on the rapid development of multi-omics technology, as well as the application of molecular genetics in the future strategies to achieve conservation and breeding of medicinal plants and efficient biosynthesis of active compounds.


Asunto(s)
Plantas Medicinales , Vías Biosintéticas , Genómica , Biología Molecular , Fitomejoramiento
8.
J Bacteriol ; 201(14)2019 07 15.
Artículo en Inglés | MEDLINE | ID: mdl-31061168

RESUMEN

Bacillus spores incubated on plates for 2 to 98 days at 37°C had identical Ca-dipicolinic acid contents, exhibited identical viability on rich- or poor-medium plates, germinated identically in liquid with all germinants tested, identically returned to vegetative growth in rich or minimal medium, and exhibited essentially identical resistance to dry heat and similar resistance to UV radiation. However, the oldest spores had a lower core water content and significantly higher wet heat and NaOCl resistance. In addition, 47- and 98-day spores had lost >98% of intact 16S and 23S rRNA and 97 to 99% of almost all mRNAs, although minimal amounts of mononucleotides were generated in 91 days. Levels of 3-phosphoglyceric acid (3PGA) also fell 30 to 60% in the oldest spores, but how the 3PGA was lost is not clear. These results indicate that (i) translation of dormant spore mRNA is not essential for completion of spore germination, nor is protein synthesis from any mRNA; (ii) in sporulation for up to 91 days at 37°C, the RNA broken down generates minimal levels of mononucleotides; and (iii) the lengths of time that spores are incubated in sporulation medium should be considered when determining conditions for spore inactivation by wet heat, in particular, in using spores to test for the efficacy of sterilization regimens.IMPORTANCE We show that spores incubated at 37°C on sporulation plates for up to 98 days have lost almost all mRNAs and rRNAs, yet the aged spores germinated and outgrew as well as 2-day spores, and all these spores had identical viability. Thus, it is unlikely that spore mRNA, rRNA, or protein synthesis is important in spore germination. Spores incubated for 47 to 98 days also had much higher wet heat resistance than 2-day spores, suggesting that spore "age" should be considered in generating spores for tests of sterilization assurance. These data are the first to show complete survival of hydrated spores for ∼100 days, complementing published data showing dry-spore survival for years.


Asunto(s)
Bacillus subtilis/crecimiento & desarrollo , Calor , Esporas Bacterianas/fisiología , Agua , Bacillus subtilis/genética , Bacillus subtilis/efectos de la radiación , Viabilidad Microbiana/efectos de la radiación , ARN Bacteriano/genética , ARN Mensajero/genética , ARN Ribosómico 16S/genética , ARN Ribosómico 23S/genética , Esporas Bacterianas/genética , Esporas Bacterianas/efectos de la radiación , Rayos Ultravioleta
9.
Appl Environ Microbiol ; 84(7)2018 04 01.
Artículo en Inglés | MEDLINE | ID: mdl-29330188

RESUMEN

DNA damage kills dry-heated spores of Bacillus subtilis, but dry-heat-treatment effects on spore germination and outgrowth have not been studied. This is important, since if dry-heat-killed spores germinate and undergo outgrowth, toxic proteins could be synthesized. Here, Raman spectroscopy and differential interference contrast microscopy were used to study germination and outgrowth of individual dry-heat-treated B. subtilis and Bacillus megaterium spores. The major findings in this work were as follows: (i) spores dry-heat-treated at 140°C for 20 min lost nearly all viability but retained their Ca2+-dipicolinic acid (CaDPA) depot; (ii) in most cases, dry-heat treatment increased the average times and variability of all major germination events in B. subtilis spore germination with nutrient germinants or CaDPA, and in one nutrient germination event with B. megaterium spores; (iii) B. subtilis spore germination with dodecylamine, which activates the spore CaDPA release channel, was unaffected by dry-heat treatment; (iv) these results indicate that dry-heat treatment likely damages spore proteins important in nutrient germinant recognition and cortex peptidoglycan hydrolysis, but not CaDPA release itself; and (v) analysis of single spores incubated on nutrient-rich agar showed that while dry-heat-treated spores that are dead can complete germination, they cannot proceed into outgrowth and thus not to vegetative growth. The results of this study provide new information on the effects of dry heat on bacterial spores and indicate that dry-heat sterilization regimens should produce spores that cannot outgrow and thus cannot synthesize potentially dangerous proteins.IMPORTANCE Much research has shown that high-temperature dry heat is a promising means for the inactivation of spores on medical devices and spacecraft decontamination. Dry heat is known to kill Bacillus subtilis spores by DNA damage. However, knowledge about the effects of dry-heat treatment on spore germination and outgrowth is limited, especially at the single spore level. In the current work, Raman spectroscopy and differential interference contrast microscopy were used to analyze CaDPA levels in and kinetics of nutrient- and non-nutrient germination of multiple individual dry-heat-treated B. subtilis and Bacillus megaterium spores that were largely dead. The outgrowth and subsequent cell division of these germinated but dead dry-heat-treated spores were also examined. The knowledge obtained in this study will help understand the effects of dry heat on spores both on Earth and in space, and indicates that dry heat can be safely used for sterilization purposes.


Asunto(s)
Bacillus megaterium/fisiología , Bacillus subtilis/fisiología , Calor , Esporas Bacterianas/fisiología , Esterilización , Cinética , Microscopía de Interferencia , Espectrometría Raman
10.
Appl Opt ; 56(12): 3263-3269, 2017 Apr 20.
Artículo en Inglés | MEDLINE | ID: mdl-28430241

RESUMEN

Bacillus thuringiensis (Bt) is the most widely used microbial insecticide. To clarify the mechanism of bacterial resistance to ethanol toxicity, the present study investigated the effects of 70% (v/v) ethanol at a moderate temperature (65°C) on Bt spore germination by single-cell Raman spectroscopy and differential interference contrast microscopy. We found that over 80% of Bt spores were inviable after 30 min of treatment. Moreover, ethanol treatment affected spore germination; the time for initiation of rapid calcium dipicolinate (CaDPA) release (i.e., lag time, Tlag), time taken for rapid CaDPA release (i.e., ΔTrelease), and time required for complete hydrolysis of the peptidoglycan cortex of spores (i.e., ΔTlys) were increased with longer treatment times. Alanine-initiated germination upon ethanol treatment for 30-90 min showed a 2- to 4-fold longer Tlag, 2- to 3.5-fold longer ΔTrelease, and ∼2-fold longer ΔTlys relative to the control. Dodecylamine-initiated germination treated for 15-30 min had 3- to 5-fold longer Tlag and 1.4- to 1.7-fold longer ΔTrelease than the control. Germination induced by exogenous CaDPA was observed only in a small fraction of spores treated with ethanol for 5 min. Single-cell Raman spectroscopy revealed that more than 52% of spores lost CaDPA after 30 min of ethanol treatment; these showed reductions in the intensity of 1280 and 1652 cm-1 bands (corresponding to protein α-helical structure) and increases in that of 1245 and 1665 cm-1 bands (attributed to irregularities in protein structure). These results indicate that CaDPA in the core of Bt spores confers resistance to ethanol, and that damage to the spore inner membrane by ethanol treatment results in CaDPA leakage. Additionally, moderate-temperature ethanol treatment and consequent denaturation of germination-related proteins affected spore germination, specifically by inactivating the cortex-lytic enzyme CwlJ. Our findings provide a theoretical basis for the development of more effective methods for killing spore-forming bacteria; microscopy imaging and Raman spectroscopy can provide novel insight into the effects of chemical agents on microbial cells.


Asunto(s)
Antiinfecciosos Locales/farmacología , Bacillus thuringiensis/efectos de los fármacos , Etanol/farmacología , Esporas Bacterianas/efectos de los fármacos , Alanina/farmacología , Aminas/farmacología , Bacillus thuringiensis/fisiología , Farmacorresistencia Bacteriana , Cinética , Microscopía de Interferencia , Ácidos Picolínicos/metabolismo , Ácidos Picolínicos/farmacología , Desnaturalización Proteica , Espectrometría Raman/métodos , Esporas Bacterianas/fisiología , Factores de Tiempo
11.
J Antimicrob Chemother ; 71(6): 1540-6, 2016 06.
Artículo en Inglés | MEDLINE | ID: mdl-26861569

RESUMEN

OBJECTIVES: Spores of Clostridium difficile and Bacillus cereus are major causes of nosocomial diarrhoea and foodborne disease. Our aim was to measure the dynamics of the uptake of the antibiotic berberine by individual germinating spores and the levels of berberine accumulated in germinated spores. METHODS: Laser tweezers Raman spectroscopy (LTRS) and differential interference contrast microscopy were used to measure levels of berberine accumulated in single germinating spores and to monitor berberine uptake and germination of individual C. difficile and B. cereus spores. RESULTS: MIC values of berberine for C. difficile and B. cereus spores were 640 and 256 mg/L, respectively. Levels of berberine accumulated at the berberine MICs in individual germinated spores were heterogeneous, with values of 17.1 ±â€Š5.4 and 12.7 ±â€Š5.5 g/L for C. difficile and B. cereus spores, respectively. These values were 25-50-fold higher than the MIC values. However, berberine did not affect the germination of C. difficile and B. cereus spores, but did block germinated spores' outgrowth. Berberine uptake kinetics were similar for these two kinds of spores. After the addition of germinants, berberine began to enter germinating spores at the time (Tlag) when rapid release of the spore core's large depot of the 1:1 chelate of Ca(2+) with dipicolinic acid began, and the level of berberine taken up was maximal shortly after spore cortex lysis was completed (Tlysis). CONCLUSIONS: LTRS can be used to measure uptake and levels of berberine in single cells. High levels of berberine can enter spores of C. difficile and B. cereus soon after germination is initiated, thus inhibiting spore outgrowth and minimizing hazards posed by germinated spores.


Asunto(s)
Antibacterianos/análisis , Bacillus cereus/química , Berberina/análisis , Clostridioides difficile/química , Pinzas Ópticas , Espectrometría Raman/métodos , Esporas Bacterianas/química , Antibacterianos/metabolismo , Bacillus cereus/efectos de los fármacos , Berberina/metabolismo , Clostridioides difficile/efectos de los fármacos , Pruebas de Sensibilidad Microbiana , Microscopía de Interferencia/métodos , Esporas Bacterianas/efectos de los fármacos
12.
Appl Environ Microbiol ; 82(19): 5775-84, 2016 10 01.
Artículo en Inglés | MEDLINE | ID: mdl-27422840

RESUMEN

UNLABELLED: Raman spectroscopy and phase-contrast microscopy were used to examine calcium dipicolinate (CaDPA) levels and rates of nutrient and nonnutrient germination of multiple individual Bacillus subtilis spores treated with cold atmospheric plasma (CAP). Major results for this work include the following: (i) >5 logs of spores deposited on glass surfaces were inactivated by CAP treatment for 3 min, while deposited spores placed inside an impermeable plastic bag were inactivated only ∼2 logs in 30 min; (ii) >80% of the spores treated for 1 to 3 min with CAP were nonculturable and retained CaDPA in their core, while >95% of spores treated with CAP for 5 to 10 min lost all CaDPA; (iii) Raman measurements of individual CAP-treated spores without CaDPA showed differences from spores that germinated with l-valine in terms of nucleic acids, lipids, and proteins; and (iv) 1 to 2 min of CAP treatment killed 99% of spores, but these spores still germinated with nutrients or exogenous CaDPA, albeit more slowly and to a lesser extent than untreated spores, while spores CAP treated for >3 min that retained CaDPA did not germinate via nutrients or CaDPA. However, even after 1 to 3 min of CAP treatment, spores germinated normally with dodecylamine. These results suggest that exposure to the present CAP configuration severely damages a spore's inner membrane and key germination proteins, such that the treated spores either lose CaDPA or can neither initiate nor complete germination with nutrients or CaDPA. Analysis of the various CAP components indicated that UV photons contributed minimally to spore inactivation, while charged particles and reactive oxygen species contributed significantly. IMPORTANCE: Much research has shown that cold atmospheric plasma (CAP) is a promising tool for the inactivation of spores in the medical and food industries. However, knowledge about the effects of plasma treatment on spore properties is limited, especially at the single-cell level. In this study, Raman spectroscopy and phase-contrast microscopy were used to analyze CaDPA levels and kinetics of nutrient- and non-nutrient-germinant-induced germination of multiple individual spores of Bacillus subtilis that were treated by a planar CAP device. The roles of different plasma species involved in spore inactivation were also investigated. The knowledge obtained in this study will aid in understanding the mechanism(s) of spore inactivation by CAP and potentially facilitate the development of more effective and efficient plasma sterilization techniques in various applications.


Asunto(s)
Bacillus subtilis/fisiología , Técnicas Bacteriológicas/métodos , Viabilidad Microbiana , Microscopía de Contraste de Fase , Ácidos Picolínicos/metabolismo , Gases em Plasma/farmacología , Espectrometría Raman , Esporas Bacterianas/fisiología , Esporas Bacterianas/química
13.
Appl Environ Microbiol ; 82(17): 5287-97, 2016 09 01.
Artículo en Inglés | MEDLINE | ID: mdl-27316969

RESUMEN

UNLABELLED: This work analyzes the high-pressure (HP) germination of spores of the food-borne pathogen Clostridium perfringens (with inner membrane [IM] germinant receptors [GRs]) and the opportunistic pathogen Clostridium difficile (with no IM GRs), which has growing implications as an emerging food safety threat. In contrast to those of spores of Bacillus species, mechanisms of HP germination of clostridial spores have not been well studied. HP treatments trigger Bacillus spore germination through spores' IM GRs at ∼150 MPa or through SpoVA channels for release of spores' dipicolinic acid (DPA) at ≥400 MPa, and DPA-less spores have lower wet heat resistance than dormant spores. We found that C. difficile spores exhibited no germination events upon 150-MPa treatment and were not heat sensitized. In contrast, 150-MPa-treated unactivated C. perfringens spores released DPA and became heat sensitive, although most spores did not complete germination by fully rehydrating the spore core, but this treatment of heat-activated spores led to almost complete germination and greater heat sensitization. Spores of both clostridial organisms released DPA during 550-MPa treatment, but C. difficile spores did not complete germination and remained heat resistant. Heat-activated 550-MPa-HP-treated C. perfringens spores germinated almost completely and became heat sensitive. However, unactivated 550-MPa-treated C. perfringens spores did not germinate completely and were less heat sensitive than spores that completed germination. Since C. difficile and C. perfringens spores use different mechanisms for sensing germinants, our results may allow refinement of HP methods for their inactivation in foods and other applications and may guide the development of commercially sterile low-acid foods. IMPORTANCE: Spores of various clostridial organisms cause human disease, sometimes due to food contamination by spores. Because of these spores' resistance to normal decontamination regimens, there is continued interest in ways to kill spores without compromising food quality. High hydrostatic pressure (HP) under appropriate conditions can inactivate bacterial spores. With growing use of HP for food pasteurization, advancement of HP for commercial production of sterile low-acid foods requires understanding of mechanisms of spores' interactions with HP. While much is known about HP germination and inactivation of spores of Bacillus species, how HP germinates and inactivates clostridial spores is less well understood. In this work we have tried to remedy this information deficit by examining germination of spores of Clostridium difficile and Clostridium perfringens by several HP and temperature levels. The results may give insight that could facilitate more efficient methods for spore eradication in food sterilization or pasteurization, biodecontamination, and health care.


Asunto(s)
Clostridium/crecimiento & desarrollo , Desinfección/métodos , Esporas Bacterianas/química , Proteínas Bacterianas/genética , Proteínas Bacterianas/metabolismo , Clostridium/química , Clostridium/genética , Clostridium/metabolismo , Presión , Esporas Bacterianas/crecimiento & desarrollo , Esporas Bacterianas/metabolismo , Temperatura
14.
Sensors (Basel) ; 16(8)2016 Aug 08.
Artículo en Inglés | MEDLINE | ID: mdl-27509509

RESUMEN

Poly-beta-hydroxybutyrate (PHB) can be formed in large amounts in Cupriavidus necator and is important for the industrial production of biodegradable plastics. In this investigation, laser tweezers Raman spectroscopy (LTRS) was used to characterize dynamic changes in PHB content-as well as in the contents of other common biomolecule-in C. necator during batch growth at both the population and single-cell levels. PHB accumulation began in the early stages of bacterial growth, and the maximum PHB production rate occurred in the early and middle exponential phases. The active biosynthesis of DNA, RNA, and proteins occurred in the lag and early exponential phases, whereas the levels of these molecules decreased continuously during the remaining fermentation process until the minimum values were reached. The PHB content inside single cells was relatively homogenous in the middle stage of fermentation; during the late growth stage, the variation in PHB levels between cells increased. In addition, bacterial cells in various growth phases could be clearly discriminated when principle component analysis was performed on the spectral data. These results suggest that LTRS is a valuable single-cell analysis tool that can provide more comprehensive information about the physiological state of a growing microbial population.


Asunto(s)
Cupriavidus necator/genética , Hidroxibutiratos/química , Cinética , Poliésteres/química , Espectrometría Raman/métodos , Plásticos Biodegradables/química , Cupriavidus necator/química , Cupriavidus necator/crecimiento & desarrollo , ADN/biosíntesis , Pinzas Ópticas , Biosíntesis de Proteínas/genética , ARN/biosíntesis , Análisis de la Célula Individual/métodos
15.
J Bacteriol ; 197(6): 1095-103, 2015 Mar.
Artículo en Inglés | MEDLINE | ID: mdl-25583976

RESUMEN

When exposed to nutrient or nonnutrient germinants, individual Bacillus spores can return to life through germination followed by outgrowth. Laser tweezers, Raman spectroscopy, and either differential interference contrast or phase-contrast microscopy were used to analyze the slow dipicolinic acid (DPA) leakage (normally ∼20% of spore DPA) from individual spores that takes place prior to the lag time, Tlag, when spores begin rapid release of remaining DPA. Major conclusions from this work with Bacillus subtilis spores were as follows: (i) slow DPA leakage from wild-type spores germinating with nutrients did not begin immediately after nutrient exposure but only at a later heterogeneous time T1; (ii) the period of slow DPA leakage (ΔTleakage = Tlag - T1) was heterogeneous among individual spores, although the amount of DPA released in this period was relatively constant; (iii) increases in germination temperature significantly decreased T1 times but increased values of ΔTleakage; (iv) upon germination with l-valine for 10 min followed by addition of d-alanine to block further germination, all germinated spores had T1 times of less than 10 min, suggesting that T1 is the time when spores become committed to germinate; (v) elevated levels of SpoVA proteins involved in DPA movement in spore germination decreased T1 and Tlag times but not the amount of DPA released in ΔTleakage; (vi) lack of the cortex-lytic enzyme CwlJ increased DPA leakage during germination due to longer ΔTleakage times in which more DPA was released; and (vii) there was slow DPA leakage early in germination of B. subtilis spores by the nonnutrients CaDPA and dodecylamine and in nutrient germination of Bacillus cereus and Bacillus megaterium spores. Overall, these findings have identified and characterized a new early event in Bacillus spore germination.


Asunto(s)
Bacillus subtilis/fisiología , Ácidos Picolínicos/metabolismo , Esporas Bacterianas/fisiología , Proteínas Bacterianas/genética , Proteínas Bacterianas/metabolismo , Regulación Bacteriana de la Expresión Génica , Temperatura , Factores de Tiempo
16.
J Bacteriol ; 197(14): 2361-73, 2015 Jul.
Artículo en Inglés | MEDLINE | ID: mdl-25939833

RESUMEN

UNLABELLED: The Gram-positive spore-forming anaerobe Clostridium difficile is a leading cause of nosocomial diarrhea. Spores of C. difficile initiate infection when triggered to germinate by bile salts in the gastrointestinal tract. We analyzed germination kinetics of individual C. difficile spores using Raman spectroscopy and differential interference contrast (DIC) microscopy. Similar to Bacillus spores, individual C. difficile spores germinating with taurocholate plus glycine began slow leakage of a ∼15% concentration of a chelate of Ca(2+) and dipicolinic acid (CaDPA) at a heterogeneous time T1, rapidly released CaDPA at Tlag, completed CaDPA release at Trelease, and finished peptidoglycan cortex hydrolysis at Tlysis. T1 and Tlag values for individual spores were heterogeneous, but ΔTrelease periods (Trelease - Tlag) were relatively constant. In contrast to Bacillus spores, heat treatment did not stimulate spore germination in the two C. difficile strains tested. C. difficile spores did not germinate with taurocholate or glycine alone, and different bile salts differentially promoted spore germination, with taurocholate and taurodeoxycholate being best. Transient exposure of spores to taurocholate plus glycine was sufficient to commit individual spores to germinate. C. difficile spores did not germinate with CaDPA, in contrast to B. subtilis and C. perfringens spores. However, the detergent dodecylamine induced C. difficile spore germination, and rates were increased by spore coat removal although cortex hydrolysis did not follow Trelease, in contrast with B. subtilis. C. difficile spores lacking the cortex-lytic enzyme, SleC, germinated extremely poorly, and cortex hydrolysis was not observed in the few sleC spores that partially germinated. Overall, these findings indicate that C. difficile and B. subtilis spore germination exhibit key differences. IMPORTANCE: Spores of the Gram-positive anaerobe Clostridium difficile are responsible for initiating infection by this important nosocomial pathogen. When exposed to germinants such as bile salts, C. difficile spores return to life through germination in the gastrointestinal tract and cause disease, but their germination has been studied only with population-wide measurements. In this work we used Raman spectroscopy and DIC microscopy to monitor the kinetics of germination of individual C. difficile spores, the commitment of spores to germination, and the effect of germinant type and concentration, sublethal heat shock, and spore decoating on germination. Our data suggest that the order of germination events in C. difficile spores differs from that in Bacillus spores and provide new insights into C. difficile spore germination.


Asunto(s)
Clostridioides difficile/fisiología , Microscopía de Interferencia/métodos , Espectrometría Raman/métodos , Esporas Bacterianas/fisiología , Calcio/química , Glicina , Calor , Ácidos Picolínicos/química , Ácido Taurocólico , Factores de Tiempo
17.
J Bacteriol ; 196(13): 2443-54, 2014 Jul.
Artículo en Inglés | MEDLINE | ID: mdl-24769693

RESUMEN

Short exposures of Bacillus spores to nutrient germinants can commit spores to germinate when germinants are removed or their binding to the spores' nutrient germinant receptors (GRs) is inhibited. Bacillus subtilis spores were exposed to germinants for various periods, followed by germinant removal to prevent further commitment. Release of spore dipicolinic acid (DPA) was then measured by differential interference contrast microscopy to monitor germination of multiple individual spores, and spores did not release DPA after 1 to 2 min of germinant exposure until ~7 min after germinant removal. With longer germinant exposures, percentages of committed spores with times for completion of DPA release (Trelease) greater than the time of germinant removal (Tb) increased, while the time Tlag - Tb, where Tlag represents the time when rapid DPA release began, was decreased but rapid DPA release times (ΔTrelease = Trelease - Tlag) were increased; Factors affecting average Trelease values and the percentages of committed spores were germinant exposure time, germinant concentration, sporulation conditions, and spore heat activation, as previously shown for commitment of spore populations. Surprisingly, germination of spores given a 2nd short germinant exposure 30 to 45 min after a 1st exposure of the same duration was significantly higher than after the 1st exposure, but the number of spores that germinated in the 2nd germinant exposure decreased as the interval between germinant exposures increased up to 12 h. The latter results indicate that spores have some memory, albeit transient, of their previous exposure to nutrient germinants.


Asunto(s)
Bacillus subtilis/fisiología , Alanina/farmacología , Asparagina/farmacología , Bacillus subtilis/efectos de los fármacos , Ácidos Picolínicos/metabolismo , Esporas Bacterianas/efectos de los fármacos , Esporas Bacterianas/metabolismo , Esporas Bacterianas/fisiología , Valina/farmacología
18.
J Bacteriol ; 196(11): 2077-88, 2014 Jun.
Artículo en Inglés | MEDLINE | ID: mdl-24682327

RESUMEN

The Bacillus subtilis spoVAEa and spoVAF genes are expressed in developing spores as members of the spoVA operon, which encodes proteins essential for the uptake and release of dipicolinic acid (DPA) during spore formation and germination. SpoVAF is likely an integral inner spore membrane protein and exhibits sequence identity to A subunits of the spore's nutrient germinant receptors (GRs), while SpoVAEa is a soluble protein with no obvious signals to allow its passage across a membrane. However, like SpoVAD, SpoVAEa is present on the outer surface of the spore's inner membrane, as SpoVAEa was accessible to an external biotinylation agent in spores and SpoVAEa disappeared in parallel with SpoVAD during proteinase K treatment of germinated spores. SpoVAEa and SpoVAD were also distributed similarly in fractions of disrupted dormant spores. Unlike spoVAD, spoVAEa is absent from the genomes of some spore-forming members of the Bacillales and Clostridiales orders, although SpoVAEa's amino acid sequence is conserved in species containing spoVAEa. B. subtilis strains lacking SpoVAF or SpoVAEa and SpoVAF sporulated normally, and the spores had normal DPA levels. Spores lacking SpoVAF or SpoVAEa and SpoVAF also germinated normally with non-GR-dependent germinants but more slowly than wild-type spores with GR-dependent germinants, and this germination defect was complemented by ectopic expression of the missing proteins.


Asunto(s)
Bacillus subtilis/fisiología , Proteínas Bacterianas/metabolismo , Regulación Bacteriana de la Expresión Génica/fisiología , Esporas Bacterianas/fisiología , Proteínas Bacterianas/genética , Prueba de Complementación Genética , Genotipo , Mutación
19.
Appl Environ Microbiol ; 80(4): 1314-21, 2014 Feb.
Artículo en Inglés | MEDLINE | ID: mdl-24317076

RESUMEN

The effect of high NaCl concentrations on nutrient and nonnutrient germination of Bacillus subtilis spores was systematically investigated. Under all conditions, increasing NaCl concentrations caused increasing, albeit reversible, inhibition of germination. High salinity delayed and increased the heterogeneity of germination initiation, slowed the germination kinetics of individual spores and the whole spore population, and decreased the overall germination efficiency, as observed by a variety of different analytical techniques. Germination triggered by nutrients which interact with different germinant receptors (GRs) was affected differently by NaCl, suggesting that GRs are targets of NaCl inhibition. However, NaCl also inhibited GR-independent germination, suggesting that there is at least one additional target for NaCl inhibition. Strikingly, a portion of the spore population could initiate germination with l-alanine even at NaCl concentrations near saturation (∼5.4 M), suggesting that spores lack a salt-sensing system preventing them from germinating in a hostile high-salinity environment. Spores that initiated germination at very high NaCl concentrations excreted their large depot of Ca(2+)-pyridine-2,6-dicarboxylic acid and lost their heat resistance, but they remained in a phase-gray state in the phase-contrast microscope, suggesting that there was incomplete germination. However, some metabolic activity could be detected at up to 4.8 M NaCl. Overall, high salinity seems to exert complex effects on spore germination and outgrowth whose detailed elucidation in future investigations could give valuable insights on these processes in general.


Asunto(s)
Bacillus subtilis/efectos de los fármacos , Bacillus subtilis/crecimiento & desarrollo , Salinidad , Cloruro de Sodio/metabolismo , Esporas Bacterianas/efectos de los fármacos , Esporas Bacterianas/crecimiento & desarrollo , Alanina/metabolismo , Bacillus subtilis/citología , Microscopía de Contraste de Fase , Ácidos Picolínicos/metabolismo , Esporas Bacterianas/citología
20.
Appl Environ Microbiol ; 80(15): 4606-15, 2014 Aug.
Artículo en Inglés | MEDLINE | ID: mdl-24837375

RESUMEN

High-precision measurements of size changes of individual bacterial spores based on ellipse fitting to bright-field images recorded with a digital camera were employed to monitor the germination of Bacillus spores with a precision of ~5 nm. To characterize the germination of individual spores, we recorded bright-field and phase-contrast images and found that the timing of changes in their normalized intensities coincided, so the bright-field images can be used to characterize spore size and refractility changes during germination. The major conclusions from this work were as follows. (i) The sizes of germinating B. cereus spores were nearly unchanged until Trelease, the time of the completion of CaDPA (a 1:1 chelate of Ca(2+) and dipicolinic acid [DPA]) release after addition of nutrient germinants. (ii) The minor axis of germinating B. cereus spores rapidly increased by ~50 nm in a few seconds right after Trelease, while the major axis was slightly decreased or unchanged. Both the minor and major axes remained unchanged for a further 30 to 45 s and then increased by 100 to 200 nm by Tlys, the time of completion of cortex lysis. (iii) Individual spores in a population showed significant heterogeneity in the timing of germination events, such as Trelease and Tlys, but also variation in size changes during germination. (iv) Bacillus subtilis wild-type spores, B. subtilis spores lacking the cortex-lytic enzyme CwlJ, and wild-type Bacillus megaterium spores showed similar kinetics of size changes during nutrient germination. The size increases in germinating spores probably result from uptake of water and cortex lysis after completion of CaDPA release.


Asunto(s)
Bacillus subtilis/crecimiento & desarrollo , Esporas Bacterianas/química , Esporas Bacterianas/citología , Bacillus subtilis/química , Bacillus subtilis/metabolismo , Calcio/metabolismo , Cinética , Ácidos Picolínicos/metabolismo , Esporas Bacterianas/metabolismo
SELECCIÓN DE REFERENCIAS
DETALLE DE LA BÚSQUEDA