[go: up one dir, main page]

Next Issue
Volume 7, June
Previous Issue
Volume 6, December
 
 

Antibiotics, Volume 7, Issue 1 (March 2018) – 26 articles

  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
11 pages, 1251 KiB  
Article
Geographic Variation in Antibiotic Consumption—Is It Due to Doctors’ Prescribing or Patients’ Consulting?
by Marte Meyer Walle-Hansen and Sigurd Høye
Antibiotics 2018, 7(1), 26; https://doi.org/10.3390/antibiotics7010026 - 20 Mar 2018
Cited by 7 | Viewed by 7028
Abstract
Antibiotic consumption varies greatly between Norwegian municipalities. We examine whether this variation is associated with inhabitants’ consultation rates or general practitioners’ (GP) prescription rates. Our study comprises consultations and antibiotic prescriptions for respiratory tract infections (RTIs) in general practice in all Norwegian municipalities [...] Read more.
Antibiotic consumption varies greatly between Norwegian municipalities. We examine whether this variation is associated with inhabitants’ consultation rates or general practitioners’ (GP) prescription rates. Our study comprises consultations and antibiotic prescriptions for respiratory tract infections (RTIs) in general practice in all Norwegian municipalities with over 5000 inhabitants in 2014. Data was collected from The Norwegian Prescription Database, The Directorate of Health’s system for control and payment of health reimbursements registry and Norway Statistics. Consultation rates and prescription rates were categorised in age- and gender specific quintiles and the effect on antibiotic consumption was analysed using a Poisson regression model. We found that inhabitants with RTIs received 42% more prescriptions if they belonged to a municipality with high consultation rates compared to low consultation rates [incidence rate ratio (IRR) 1.42 (95% CI 1.41–1.44)] and 48% more prescriptions if they belonged to a municipality with high prescription rates versus low prescription rates [IRR 1.48 (95% KI 1.47–1.50)]. Our results demonstrate that inhabitants’ consultation rates and GPs’ prescription rates have about equal impact on the number of RTI antibiotics prescribed at municipality level. These findings highlight the importance of interventions targeting patients as well as doctors in efforts to reduce unnecessary antibiotic consumption. Full article
Show Figures

Figure 1

Figure 1
<p>RTI antibiotic prescriptions per 1000 inhabitants in 198 Norwegian municipalities. Quintile 1: 76–186 prescriptions. Quintile 2: 186–209 prescriptions. Quintile 3: 209–224 prescriptions. Quintile 4: 224–246 prescriptions. Quintile 5: 246–331 prescriptions [<a href="#B13-antibiotics-07-00026" class="html-bibr">13</a>].</p>
Full article ">Figure 2
<p>RTI consultations per 1000 inhabitants in 198 Norwegian municipalities. Quintile 1: 158–263 consultations. Quintile 2: 266–293 consultations. Quintile 3: 294–314 consultations. Quintile 4: 315–344 consultations. Quintile 5: 344–412 consultations [<a href="#B13-antibiotics-07-00026" class="html-bibr">13</a>].</p>
Full article ">Figure 3
<p>RTI antibiotic prescriptions per 1000 RTI consultations in 198 Norwegian municipalities. Quintile 1: 271–634 prescriptions per 1000 RTI consultations. Quintile 2: 636–688 prescriptions per 1000 RTI consultations. Quintile 3: 688–736 prescriptions per 1000 RTI consultations. Quintile 4: 738–792 prescriptions per 1000 RTI consultations. Quintile 5: 793–1146 prescriptions per 1000 RTI consultations [<a href="#B13-antibiotics-07-00026" class="html-bibr">13</a>].</p>
Full article ">
13 pages, 2658 KiB  
Review
Novel Aspects of Polynucleotide Phosphorylase Function in Streptomyces
by George H. Jones
Antibiotics 2018, 7(1), 25; https://doi.org/10.3390/antibiotics7010025 - 18 Mar 2018
Cited by 6 | Viewed by 4671
Abstract
Polynucleotide phosphorylase (PNPase) is a 3′–5′-exoribnuclease that is found in most bacteria and in some eukaryotic organelles. The enzyme plays a key role in RNA decay in these systems. PNPase structure and function have been studied extensively in Escherichia coli, but there [...] Read more.
Polynucleotide phosphorylase (PNPase) is a 3′–5′-exoribnuclease that is found in most bacteria and in some eukaryotic organelles. The enzyme plays a key role in RNA decay in these systems. PNPase structure and function have been studied extensively in Escherichia coli, but there are several important aspects of PNPase function in Streptomyces that differ from what is observed in E. coli and other bacterial genera. This review highlights several of those differences: (1) the organization and expression of the PNPase gene in Streptomyces; (2) the possible function of PNPase as an RNA 3′-polyribonucleotide polymerase in Streptomyces; (3) the function of PNPase as both an exoribonuclease and as an RNA 3′-polyribonucleotide polymerase in Streptomyces; (4) the function of (p)ppGpp as a PNPase effector in Streptomyces. The review concludes with a consideration of a number of unanswered questions regarding the function of Streptomyces PNPase, which can be examined experimentally. Full article
(This article belongs to the Special Issue Actinomycetes: The Antibiotics Producers)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the <span class="html-italic">Streptomyces coelicolor rpsO–pnp</span> operon. P<span class="html-italic">rpsO</span>A, B and P<span class="html-italic">pnp</span>A, B represent the upstream and intergenic promoters found in <span class="html-italic">S. coelicolor</span>, respectively. The ball-and-stick structures immediately following <span class="html-italic">rpsO</span> and <span class="html-italic">pnp</span> represent rho-independent transcription terminators. The ball-and-stick structure just upstream of <span class="html-italic">pnp</span> represents the intergenic hairpin which is cleaved by RNase III. The diagram is not drawn to scale.</p>
Full article ">Figure 2
<p>(<b>A</b>) Growth of the <span class="html-italic">S. coelicolor</span> strains containing promoter probe constructs. Growth was measured as the increase in optical density at 450 nm. The arrows in the figure indicate the onset of the production of two of the secondary metabolites synthesized by <span class="html-italic">S. coelicolor</span>, undecylprodigiosin (red) and actinorhodin (act). (<b>B</b>) Catechol dioxygenase (CATO<sub>2</sub>ase) activity of mycelial extracts of <span class="html-italic">S. coelicolor</span> derivatives containing the putative <span class="html-italic">rpsO</span>A and <span class="html-italic">rpsO</span>B promoters, cloned in the promoter probe vector pIPP2 [<a href="#B35-antibiotics-07-00025" class="html-bibr">35</a>]. Mycelium was harvested at the indicated times, disrupted by sonication, and following centrifugation, supernatants were assayed for catechol dioxygenase, as described previously [<a href="#B31-antibiotics-07-00025" class="html-bibr">31</a>,<a href="#B35-antibiotics-07-00025" class="html-bibr">35</a>]. The catechol dioxygenase gene is the reporter in the promoter probe vector [<a href="#B35-antibiotics-07-00025" class="html-bibr">35</a>]. (<b>C</b>) CATO<sub>2</sub>ase activities of extracts of strains containing the putative <span class="html-italic">pnp</span>A and <span class="html-italic">pnp</span>B promoters. The results shown are the averages of duplicate assays from two independent experiments ± SEM. This figure is reprinted from <span class="html-italic">Gene</span>, 536, Patricia Bralley, Marcha L. Gatewood, George H. Jones, Transcription of the <span class="html-italic">rpsO–pnp</span> operon of <span class="html-italic">Streptomyces coelicolor</span> involves four temporally regulated, stress responsive promoters. 177–185, Copyright (2014), with permission from Elsevier [<a href="#B31-antibiotics-07-00025" class="html-bibr">31</a>].</p>
Full article ">Figure 3
<p>Cold shock responses of <span class="html-italic">S. coelicolor</span>. Derivatives containing the <span class="html-italic">rpsO–pnp</span> promoter probe constructs were grown and 30 °C, and half of each culture was then shifted to 10 °C. Mycelium was harvested at the indicated times, disrupted by sonication, and following centrifugation, supernatants were assayed for promoter activity, as described [<a href="#B31-antibiotics-07-00025" class="html-bibr">31</a>,<a href="#B35-antibiotics-07-00025" class="html-bibr">35</a>]. Panel <b>C</b> shows the results of PNPase polymerization assays. In Panels <b>A</b> and <b>B</b>, PNPase promoter activities are expressed relative to the activity measured at 30 °C at zero time, immediately before the shift to 10 °C. The results shown are the averages of duplicate assays from two independent experiments ± SEM. In the first experiment, PNPase levels were measured in <span class="html-italic">S. coelicolor</span> containing P<span class="html-italic">rpsO</span>A and in the second, PNPase levels were measured in the derivative containing P<span class="html-italic">pnp</span>B. This figure is reprinted from Gene, 536, Patricia Bralley, Marcha L. Gatewood, George H. Jones, Transcription of the <span class="html-italic">rpsO-pnp</span> operon of <span class="html-italic">Streptomyces coelicolor</span> involves four temporally regulated, stress responsive promoters. 177–185, Copyright (2014), with permission from Elsevier [<a href="#B31-antibiotics-07-00025" class="html-bibr">31</a>].</p>
Full article ">Figure 4
<p>Effects of nucleoside diphosphates on the phosphorolysis of the 5650 transcript. Phosphorolysis reactions were performed as described in [<a href="#B47-antibiotics-07-00025" class="html-bibr">47</a>], and reaction products were separated by gel electrophoresis. The top panel shows the results obtained with <span class="html-italic">S. coelicolor</span> PNPase and the bottom panel results using <span class="html-italic">E. coli</span> PNPase. Reactions were conducted in the presence of increasing concentrations of a mixture of ADP, CDP, UDP, and GDP (nucleoside diphosphates (NDPs)) as indicated. RP3 is the 5650 transcript, and RP4 represents the product obtained by complete digestion of the intergenic hairpin in RP3 by PNPase. Note that as PNPase is highly processive [<a href="#B48-antibiotics-07-00025" class="html-bibr">48</a>], no intermediates with mobilities between those of RP3 and RP4 were observed. Copyright © American Society for Microbiology (<span class="html-italic">J. Bacteriol.</span> 190, 2008, 98–106, DOI:10.1128/JB.00327-07) [<a href="#B47-antibiotics-07-00025" class="html-bibr">47</a>].</p>
Full article ">Figure 5
<p>Model for the effects of NDPs on the activity <span class="html-italic">S. coelicolor</span> PNPase. The model posits that <span class="html-italic">S. coelicolor</span> PNPase (PacMan symbol) is able to phosphorolyze 5650 and other structured substrates to a limited extent in the absence of NDPs, as indicated by the dashed X. In the presence of NDPs, PNPase synthesizes unstructured 3′-tails in vivo, and these tails then provide an anchor for the enzyme, thus facilitating the digestion of structured substrates. Copyright © American Society for Microbiology (<span class="html-italic">J. Bacteriol.</span> 195, 2013, 5151–5159, DOI:10.1128/JB.00936-13) [<a href="#B49-antibiotics-07-00025" class="html-bibr">49</a>].</p>
Full article ">Figure 6
<p>Effects of (p)ppGpp on the activity of PNPase. Polymerization and phosphorolysis reactions were performed in the absence and presence of guanosine tetraphosphate (ppGpp) or guanosine pentaphosphate (pppGpp), using purified PNPase from <span class="html-italic">S. coelicolor</span> and <span class="html-italic">E. coli</span> [<a href="#B59-antibiotics-07-00025" class="html-bibr">59</a>]. Results are expressed relative to the activities measured in the absence of (p)ppGpp, set arbitrarily to 100%.</p>
Full article ">
9 pages, 923 KiB  
Article
Parallel Colorimetric Quantification of Choline and Phosphocholine as a Method for Studying Choline Kinase Activity in Complex Mixtures
by Tahl Zimmerman and Salam A. Ibrahim
Antibiotics 2018, 7(1), 24; https://doi.org/10.3390/antibiotics7010024 - 17 Mar 2018
Cited by 8 | Viewed by 4977
Abstract
Choline kinase (Chok) is an enzyme found in eukaryotes and Gram-positive bacteria. Chok catalyzes the production of phosphocholine from choline and ATP. This enzyme has been validated as a drug target in Streptococcus pneumonia, but the role Chok enzymatic activity plays in [...] Read more.
Choline kinase (Chok) is an enzyme found in eukaryotes and Gram-positive bacteria. Chok catalyzes the production of phosphocholine from choline and ATP. This enzyme has been validated as a drug target in Streptococcus pneumonia, but the role Chok enzymatic activity plays in bacterial cell growth and division is not well understood. Phosphocholine production by Chok and its attenuation by inhibitors in the context of complex samples such as cell extracts can currently be quantified by several methods. These include choline depletion measurements, radioactive methods, mass-spectrometry, and nuclear magnetic resonance. The first does not measure phosphocholine directly, the second requires elaborate safety procedures, and the third and fourth require significant capital investments and technical expertise. For these reasons, a less expensive, higher throughput, more easily accessible assay is needed to facilitate further study in Gram-positive Choks. Here, we present the development of a triiodide/activated charcoal/molybdenum blue system for detecting and quantifying choline and phosphocholine in parallel. We demonstrate that this system can reliably quantify changes in choline and phosphocholine concentrations over time in Chok enzymatic assays using cell extracts as the source of the enzyme. This is an easily accessible, convenient, robust, and economical method for studying Chok activity in complex samples. The triiodide/activated charcoal/molybdenum blue system opens new doors into the study choline kinase in Gram-positive pathogens. Full article
(This article belongs to the Special Issue Top 35 of Antibiotics Travel Awards 2017)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Validation of the colorimetric method. A 1 mM concentration of each reagent was assayed, except for 2.7 mM of HC-3. (<b>A</b>) Triiodide reactions with each component of the Chok reaction: choline (Cho), ATP, and phopshocholine (PCho). (<b>B</b>) Analysis of MBD absorbance measurements after processing of each compound with triiodide step alone and the combined steps of triiodide and charcoal (marked with an (F)). (<b>C</b>) Absorbance values of PCho and Cho samples detected with and without processing (P). (<b>D</b>) SDS-PAGE of extracts of uninduced (1) and induced (2) BL21 (DE3) cells transformed with <span class="html-italic">S. pneumoniae LicA</span>, the gene coding for sChok.</p>
Full article ">Figure 2
<p>Standard concentration vs absorbance curves of Cho and PCho and colorimetric absorbance at different time points in an sChok enzymatic assay. (<b>A</b>) Standard curve of [Cho] vs absorbance derived from known quantities of Cho. (<b>B</b>) Standard curve of [PCho] vs. absorbance derived from known quantities of PCho. (<b>C</b>) Absorbance changes over time derived triiodide analysis of an sChok enzymatic assay (Control Reaction). (<b>D</b>) Absorbance changes over time derived from MBD analysis of an sChok enzymatic assay with (HC-3 reaction) and without (Control Reaction) sChok inhibitor HC-3.</p>
Full article ">
28 pages, 1054 KiB  
Review
The Macromolecular Machines that Duplicate the Escherichia coli Chromosome as Targets for Drug Discovery
by Jon M. Kaguni
Antibiotics 2018, 7(1), 23; https://doi.org/10.3390/antibiotics7010023 - 14 Mar 2018
Cited by 17 | Viewed by 7197
Abstract
DNA replication is an essential process. Although the fundamental strategies to duplicate chromosomes are similar in all free-living organisms, the enzymes of the three domains of life that perform similar functions in DNA replication differ in amino acid sequence and their three-dimensional structures. [...] Read more.
DNA replication is an essential process. Although the fundamental strategies to duplicate chromosomes are similar in all free-living organisms, the enzymes of the three domains of life that perform similar functions in DNA replication differ in amino acid sequence and their three-dimensional structures. Moreover, the respective proteins generally utilize different enzymatic mechanisms. Hence, the replication proteins that are highly conserved among bacterial species are attractive targets to develop novel antibiotics as the compounds are unlikely to demonstrate off-target effects. For those proteins that differ among bacteria, compounds that are species-specific may be found. Escherichia coli has been developed as a model system to study DNA replication, serving as a benchmark for comparison. This review summarizes the functions of individual E. coli proteins, and the compounds that inhibit them. Full article
(This article belongs to the Special Issue Bacterial DNA Replication and Replication Inhibitors)
Show Figures

Figure 1

Figure 1
<p>Replication initiation at the <span class="html-italic">E. coli</span> chromosomal origin involves the recruitment of DnaA, DnaB and DnaC to form the prepriming complex, followed by activation of DnaB, primer formation by primase, and DNA replication by DNA polymerase III holoenzyme. Shown at the top, the replication origin (<span class="html-italic">oriC</span>) of <span class="html-italic">E. coli</span> contains binding sites for Fis and IHF, and the DnaA boxes named R1-R5 that are recognized by DnaA in which the ATP or ADP bound to DnaA may affect the affinities to the respective sites [<a href="#B17-antibiotics-07-00023" class="html-bibr">17</a>,<a href="#B19-antibiotics-07-00023" class="html-bibr">19</a>,<a href="#B90-antibiotics-07-00023" class="html-bibr">90</a>,<a href="#B91-antibiotics-07-00023" class="html-bibr">91</a>]. In contrast, DnaA-ATP and not DnaA-ADP specifically binds to I-, τ- and C-sites. The sites named C3 and C2 overlap R3 and may be separate sites or part of R3 [<a href="#B17-antibiotics-07-00023" class="html-bibr">17</a>,<a href="#B19-antibiotics-07-00023" class="html-bibr">19</a>,<a href="#B90-antibiotics-07-00023" class="html-bibr">90</a>,<a href="#B91-antibiotics-07-00023" class="html-bibr">91</a>]. (<b>1</b>) DnaA, which has four functional domains as noted in the figure of DnaA, recognizes specific DNA sites in <span class="html-italic">E. coli oriC</span> to form a DnaA oligomer. DnaA then unwinds a region containing the 13mers named L, M and R; (<b>2</b>) Domain I of DnaA interacts with the N-terminal domain of DnaB in the DnaB-DnaC complex to load the complex onto the top and bottom DNA strands of the unwound region, forming a macromolecular entity named the prepriming complex. The shaded rectangle represents the space between adjacent DnaB protomers through which the single-stranded DNA passes during helicase loading; (<b>3</b>) Primase interacts with the N-terminal domain of DnaB, which is required for primer synthesis. In the transition to the next step, the open space between adjacent DnaB protomers presumably closes; (<b>4</b>) Primer synthesis (shown as red wavy lines) by primase on the top and bottom strands and the translocation of DnaB leads to the dissociation of DnaC from DnaB; (<b>5</b>) After primer synthesis, primase will dissociate from DnaB as the primer is bound by DNA polymerase III holoenzyme. DnaB will move to the junction of each replication fork; (<b>6</b>) DNA polymerase III holoenzyme extends the primers for the synthesis of each leading strand. DnaB at the junction of each replication fork unwinds the parental duplex DNA. The transient interaction of DnaB with primase as the helicase moves leads to the synthesis of subsequent primers that are extended by DNA polymerase III holoenzyme in the synthesis of Okazaki fragments. The dashed lines represent the contacts between two units of DNA polymerase III holoenzyme, forming a dimer in the coordinated synthesis of the leading and lagging strands.</p>
Full article ">Figure 2
<p>Subunit composition and their interactions of DNA polymerase III holoenzyme (reviewed in [<a href="#B207-antibiotics-07-00023" class="html-bibr">207</a>,<a href="#B218-antibiotics-07-00023" class="html-bibr">218</a>,<a href="#B224-antibiotics-07-00023" class="html-bibr">224</a>]). The subassemblies of DNA polymerase III holoenzyme are the sliding clamp composed of two DnaN or β subunits, the clamp loader or DnaX complex composed of seven subunits, and DNA polymerase III core containing the α, ε and θ subunits. The diagram also summarizes how these subunits interact within each subassembly and between subassemblies.</p>
Full article ">
11 pages, 491 KiB  
Article
Potentially Important Therapeutic Interactions between Antibiotics, and a Specially Engineered Emulsion Drug Vehicle Containing Krill-Oil-Based Phospholipids and Omega-3 Fatty Acids
by David F. Driscoll
Antibiotics 2018, 7(1), 22; https://doi.org/10.3390/antibiotics7010022 - 10 Mar 2018
Cited by 2 | Viewed by 4271
Abstract
The incidence of antimicrobial resistance (AMR) worldwide is increasing as the pipeline for the development of new chemotherapeutic entities is decreasing. Clearly, overexposure to antibiotics, including excessive dosing, is a key factor that fuels AMR. In fact, most of the new antibacterial agents [...] Read more.
The incidence of antimicrobial resistance (AMR) worldwide is increasing as the pipeline for the development of new chemotherapeutic entities is decreasing. Clearly, overexposure to antibiotics, including excessive dosing, is a key factor that fuels AMR. In fact, most of the new antibacterial agents under development are derivatives of existing classes of antibiotics. Novel approaches involving unique antimicrobial combinations, targets, and/or delivery systems are under intense investigation. An innovative combination of active pharmaceutical ingredients (APIs) consisting of antimicrobial drug(s), krill-oil-based phospholipids, and omega-3 fatty acid triglycerides, that may extend the therapeutic viability of currently effective antibiotics, at least until new chemical entities are introduced, is described. Full article
Show Figures

Figure 1

Figure 1
<p>Triacylglycerol backbone and polar carbonyl groups.</p>
Full article ">
13 pages, 1759 KiB  
Article
Protein Expression Modifications in Phage-Resistant Mutants of Aeromonas salmonicida after AS-A Phage Treatment
by Catarina Moreirinha, Nádia Osório, Carla Pereira, Sara Simões, Ivonne Delgadillo and Adelaide Almeida
Antibiotics 2018, 7(1), 21; https://doi.org/10.3390/antibiotics7010021 - 8 Mar 2018
Cited by 6 | Viewed by 4628
Abstract
The occurrence of infections by pathogenic bacteria is one of the main sources of financial loss for the aquaculture industry. This problem often cannot be solved with antibiotic treatment or vaccination. Phage therapy seems to be an alternative environmentally-friendly strategy to control infections. [...] Read more.
The occurrence of infections by pathogenic bacteria is one of the main sources of financial loss for the aquaculture industry. This problem often cannot be solved with antibiotic treatment or vaccination. Phage therapy seems to be an alternative environmentally-friendly strategy to control infections. Recognizing the cellular modifications that bacteriophage therapy may cause to the host is essential in order to confirm microbial inactivation, while understanding the mechanisms that drive the development of phage-resistant strains. The aim of this work was to detect cellular modifications that occur after phage AS-A treatment in A. salmonicida, an important fish pathogen. Phage-resistant and susceptible cells were subjected to five successive streak-plating steps and analysed with infrared spectroscopy, a fast and powerful tool for cell study. The spectral differences of both populations were investigated and compared with a phage sensitivity profile, obtained through the spot test and efficiency of plating. Changes in protein associated peaks were found, and these results were corroborated by 1-D electrophoresis of intracellular proteins analysis and by phage sensitivity profiles. Phage AS-A treatment before the first streaking-plate step clearly affected the intracellular proteins expression levels of phage-resistant clones, altering the expression of distinct proteins during the subsequent five successive streak-plating steps, making these clones recover and be phenotypically more similar to the sensitive cells. Full article
(This article belongs to the Special Issue Bacteriophages: Alternatives to Antibiotics and Beyond)
Show Figures

Figure 1

Figure 1
<p>Spot test results using a phage-resistant mutant of phage AS-A and phage AS-A after first (<b>A</b>) and fifth streak-plating steps (<b>B</b>).</p>
Full article ">Figure 2
<p>Scores scatter plot of the IR spectra of phage-resistant colonies A, B and C, along the 5 streak plating steps, and control phage sensitive colonies after 1 (Ct1) and 5 (Ct5) streaking steps. The letters correspond to the different colonies (A is colony A; B is colony B; C is colony C) and the numbers to the streaking-plate days (1 is day 1; 2 is day 2; 3 is day 3; 4 is day 4; 5 is day 5).</p>
Full article ">Figure 3
<p>Loadings plot profile of PC1 corresponding to the IR spectra of the phage-resistant colonies A, B and C, along the 5 streak-plating steps, and control phage sensitive colonies.</p>
Full article ">Figure 4
<p>(<b>A</b>) SDS PAGE gel of the intracellular proteins of <span class="html-italic">A. salmonicida</span> on first streak-plating. MW, molecular weight marker; Ct, control phage-sensitive <span class="html-italic">A. salmonicida</span>; A is Colony A of the phage-resistant <span class="html-italic">A. salmonicida</span> mutant; B is Colony B of the phage-resistant <span class="html-italic">A. salmonicida</span> mutant; C is Colony C of the phage-resistant <span class="html-italic">A. salmonicida</span> mutant. The marked bands are the ones that showed differential expression between control and clones A, B and C. Band weight is expressed in kilodalton (KDa). (<b>B</b>) Differential expression of the bands, in percentage, comparing Control (phage-sensitive <span class="html-italic">A. salmonicida</span>) with clones A, B and C (phage-resistant <span class="html-italic">A. salmonicida</span>) after 1 streak-plating steps. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>(<b>A</b>) SDS PAGE gel of the intracellular proteins of <span class="html-italic">A. salmonicida</span> on fifth streak plating. MW, molecular weight marker; Ct, control phage-sensitive <span class="html-italic">A. salmonicida</span>; A is Colony A of the phage-resistant <span class="html-italic">A. salmonicida</span> mutant; B is Colony B of the phage-resistant <span class="html-italic">A. salmonicida</span> mutant; C is Colony C of the phage-resistant <span class="html-italic">A. salmonicida</span> mutant. The marked bands are the ones that showed differential expression between control and clones A, B and C. Band weight is expressed in kilodalton (KDa). (<b>B</b>) Differential expression of the bands, in percentage, comparing Control (phage-sensitive <span class="html-italic">A. salmonicida</span>) with clones A, B and C (phage-resistant <span class="html-italic">A. salmonicida</span>) after 5 streak-plating steps. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
8 pages, 1160 KiB  
Article
Biosynthesis of Rishirilide B
by Philipp Schwarzer, Julia Wunsch-Palasis, Andreas Bechthold and Thomas Paululat
Antibiotics 2018, 7(1), 20; https://doi.org/10.3390/antibiotics7010020 - 7 Mar 2018
Cited by 9 | Viewed by 6508
Abstract
Rishirilide B was isolated from Streptomyces rishiriensis and Streptomyces bottropensis on the basis of its inhibitory activity towards alpha-2-macroglobulin. The biosynthesis of rishirilide B was investigated by feeding experiments with different 13C labelled precursors using the heterologous host Streptomyces albus J1074::cos4 containing [...] Read more.
Rishirilide B was isolated from Streptomyces rishiriensis and Streptomyces bottropensis on the basis of its inhibitory activity towards alpha-2-macroglobulin. The biosynthesis of rishirilide B was investigated by feeding experiments with different 13C labelled precursors using the heterologous host Streptomyces albus J1074::cos4 containing a cosmid encoding of the gene cluster responsible for rishirilide B production. NMR spectroscopic analysis of labelled compounds demonstrate that the tricyclic backbone of rishirilide B is a polyketide synthesized from nine acetate units. One of the acetate units is decarboxylated to give a methyl group. The origin of the starter unit was determined to be isobutyrate. Full article
(This article belongs to the Special Issue Actinomycetes: The Antibiotics Producers)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structure of rishirilide B.</p>
Full article ">Figure 2
<p>Labelling positions from feeding experiments using different labelled acetates and <span class="html-small-caps">l-</span>valine.</p>
Full article ">Figure 3
<p>Proposed pathway of rishirilide B biosynthesis. (<b>a</b>) cyclisation and decarboxylation; (<b>b</b>) aromatization, oxidation; (<b>c</b>) Bayer-Villiger oxidation; (<b>d</b>) hydrolytic ring opening; (<b>e</b>) aldole condensation.</p>
Full article ">
9 pages, 2176 KiB  
Article
Genetic Determinants of Tetracycline Resistance in Clinical Streptococcus pneumoniae Serotype 1 Isolates from Niger
by Sani Ousmane, Bouli A. Diallo and Rasmata Ouedraogo
Antibiotics 2018, 7(1), 19; https://doi.org/10.3390/antibiotics7010019 - 6 Mar 2018
Cited by 9 | Viewed by 4991
Abstract
Streptococcus pneumoniae serotype 1 is the first cause of pneumococcal meningitis Niger. To determine the underlying mechanism of resistance to tetracycline in serotype 1 Streptococcus pneumoniae, a collection of 37 isolates recovered from meningitis patients over the period of 2002 to 2009 [...] Read more.
Streptococcus pneumoniae serotype 1 is the first cause of pneumococcal meningitis Niger. To determine the underlying mechanism of resistance to tetracycline in serotype 1 Streptococcus pneumoniae, a collection of 37 isolates recovered from meningitis patients over the period of 2002 to 2009 in Niger were analyzed for drug susceptibility, and whole genome sequencing (WGS) was performed for molecular analyses. MIC level was determined for 31/37 (83.8%) isolates and allowed detection of full resistance (MIC = 8 µg) in 24/31 (77.4%) isolates. No resistance was found to macrolides and quinolones. Sequence-types deduced from WGS were ST217 (54.1%), ST303 (35.1%), ST2206 (5.4%), ST2839 (2.7%) and one undetermined ST (2.7%). All tetracycline resistant isolates carried a Tn5253 like element, which was found to be an association of two smaller transposons of Tn916 and Tn5252 families. No tet(O) and tet(Q) genes were detected. However, a tet(M) like sequence was identified in all Tn5253 positive strains and was found associated to Tn916 composite. Only one isolate was phenotypically resistant to chloramphenicol, wherein a chloramphenicol acetyl transferase (cat) gene sequence homologous to catpC194 from the Staphylococcus aureus plasmid pC194 was detected. In conclusion, clinical Streptococcus pneumoniae type 1 isolated during 2002 to 2009 meningitis surveillance in Niger were fully susceptible to macrolides and quinolones but highly resistant to tetracycline (77.4%) through acquisition of a defective Tn5253 like element composed of Tn5252 and Tn916 transposons. Of the 31 tested isolates, only one was exceptionally resistant to chloramphenicol and carried a Tn5253 transposon that contained cat gene sequence. Full article
Show Figures

Figure 1

Figure 1
<p>Percent antibiotics susceptibility levels of 31 clinical serotype1 <span class="html-italic">Streptococcus pneumoniae</span> recovered from meningitis patients during 2002–2009 meningitis surveillance in Niger. Legend: S = Susceptible; R = Resistant; PENI = Penicillin; ERY= Erythromycin; CLIND = Clindamycin; CHL = Chloramphenicol; TET = Tetracycline; CIP = Ciprofloxacin.</p>
Full article ">Figure 2
<p>Annual percent distribution of tetracycline resistant <span class="html-italic">Streptococcus pneumoniae</span> serotype 1 recovered from meningitis cases from 2002 to 2009 in Niger. The figure shows that resistant isolates prevailed over susceptible ones every year except in 2009. No isolate from 2007 was tested.</p>
Full article ">Figure 3
<p>Evolutionary diversity of <span class="html-italic">tet</span>(M) genes from clinical serotype 1 <span class="html-italic">Streptococcus pneumoniae</span> recovered from meningitis cases between 2002 and 2009 in Niger. The analysis involved 32 <span class="html-italic">tet</span>(M) sequences and a reference sequence. PN34 (Accession number: AY466395.1). Sequence names were coded as NIG for Niger followed by isolate identity number, underscore and the year of isolation. Taxa fall in two clades each split into 2 subclades.</p>
Full article ">Figure 4
<p>Comparative analysis of Tn<span class="html-italic">5253</span> mobile element (accession N° gi|284803504| to Tn<span class="html-italic">5253</span> like sequence from genomes of type 1 clinical <span class="html-italic">Streptococcus pneumoniae</span> strains NIG0588_02 representing isolates susceptible to chloramphenicol and strain NIG1144_06 that was chloramphenicol resistant.</p>
Full article ">
17 pages, 1222 KiB  
Article
Efflux Activity Differentially Modulates the Levels of Isoniazid and Rifampicin Resistance among Multidrug Resistant and Monoresistant Mycobacterium tuberculosis Strains
by Diana Machado, João Perdigão, Isabel Portugal, Marco Pieroni, Pedro A. Silva, Isabel Couto and Miguel Viveiros
Antibiotics 2018, 7(1), 18; https://doi.org/10.3390/antibiotics7010018 - 3 Mar 2018
Cited by 23 | Viewed by 5095
Abstract
With the growing body of knowledge on the contribution of efflux activity to Mycobacterium tuberculosis drug resistance, increased attention has been given to the use of efflux inhibitors as adjuvants of tuberculosis therapy. Here, we investigated how efflux activity modulates the levels of [...] Read more.
With the growing body of knowledge on the contribution of efflux activity to Mycobacterium tuberculosis drug resistance, increased attention has been given to the use of efflux inhibitors as adjuvants of tuberculosis therapy. Here, we investigated how efflux activity modulates the levels of efflux between monoresistant and multi- and extensively drug resistant (M/XDR) M. tuberculosis clinical isolates. The strains were characterized by antibiotic susceptibility testing in the presence/absence of efflux inhibitors, molecular typing, and genetic analysis of drug-resistance-associated genes. Efflux activity was quantified by real-time fluorometry. The results demonstrated that all the M. tuberculosis clinical strains, susceptible or resistant, presented a faster, rapid, and non-specific efflux-mediated short-term response to drugs. The synergism assays demonstrated that the efflux inhibitors were more effective in reducing the resistance levels in the M/XDR strains than in the monoresistant strains. This indicated that M/XDR strains presented a more prolonged response to drugs mediated by efflux compared to the monoresistant strains, but both maintain it as a long-term stress response. This work shows that efflux activity modulates the levels of drug resistance between monoresistant and M/XDR M. tuberculosis clinical strains, allowing the bacteria to survive in the presence of noxious compounds. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic representation of parameters used for the characterization of the efflux activity of the <span class="html-italic">M. tuberculosis</span> strains. Presented in the Figure are the results obtained for the H37Rv strain. (<b>A</b>) SpanEtBr corresponds to the difference between the ethidium bromide fluorescence values at t<sub>0</sub> of the highest concentration tested (Y<sub>hc</sub>) and the fluorescence value at t<sub>0</sub> of the equilibrium concentration of ethidium bromide (Y<sub>eqc</sub>); (<b>B</b>) RFF (relative final fluorescence) is a measure of how effective a compound is on the inhibition of ethidium bromide efflux (at a given concentration) by comparison of the final fluorescence at the last time point (60 min) of the treated cells with the cells in presence of ethidium bromide only [<a href="#B19-antibiotics-07-00018" class="html-bibr">19</a>,<a href="#B32-antibiotics-07-00018" class="html-bibr">32</a>]; (<b>C</b>) efflux rate constant, or K value, and the t<sub>efflux50%</sub> that corresponds to the time required for the cells to extrude half of the preloaded dye; the efflux of ethidium bromide is initiated at t<sub>0</sub> by the addition of 0.4% glucose and is 50% complete at t<sub>efflux50%</sub>.</p>
Full article ">Figure 2
<p>Accumulation and efflux of ethidium bromide of the <span class="html-italic">M. tuberculosis</span> strains. (<b>A</b>) Accumulation of ethidium bromide in the presence of efflux inhibitors. In these cases, the strains were loaded with 0.25 µg/mL of ethidium bromide in the presence of verapamil or thioridazine at ½ MIC; (<b>B</b>) Efflux of ethidium bromide. Strains were loaded with ethidium bromide at 0.25 µg/mL; efflux took place in the presence of glucose which was inhibited by verapamil at ½ MIC.</p>
Full article ">
25 pages, 902 KiB  
Review
Potential for Bacteriophage Endolysins to Supplement or Replace Antibiotics in Food Production and Clinical Care
by Michael J. Love, Dinesh Bhandari, Renwick C. J. Dobson and Craig Billington
Antibiotics 2018, 7(1), 17; https://doi.org/10.3390/antibiotics7010017 - 27 Feb 2018
Cited by 121 | Viewed by 12730
Abstract
There is growing concern about the emergence of bacterial strains showing resistance to all classes of antibiotics commonly used in human medicine. Despite the broad range of available antibiotics, bacterial resistance has been identified for every antimicrobial drug developed to date. Alarmingly, there [...] Read more.
There is growing concern about the emergence of bacterial strains showing resistance to all classes of antibiotics commonly used in human medicine. Despite the broad range of available antibiotics, bacterial resistance has been identified for every antimicrobial drug developed to date. Alarmingly, there is also an increasing prevalence of multidrug-resistant bacterial strains, rendering some patients effectively untreatable. Therefore, there is an urgent need to develop alternatives to conventional antibiotics for use in the treatment of both humans and food-producing animals. Bacteriophage-encoded lytic enzymes (endolysins), which degrade the cell wall of the bacterial host to release progeny virions, are potential alternatives to antibiotics. Preliminary studies show that endolysins can disrupt the cell wall when applied exogenously, though this has so far proven more effective in Gram-positive bacteria compared with Gram-negative bacteria. Their potential for development is furthered by the prospect of bioengineering, and aided by the modular domain structure of many endolysins, which separates the binding and catalytic activities into distinct subunits. These subunits can be rearranged to create novel, chimeric enzymes with optimized functionality. Furthermore, there is evidence that the development of resistance to these enzymes may be more difficult compared with conventional antibiotics due to their targeting of highly conserved bonds. Full article
(This article belongs to the Special Issue Bacteriophages: Alternatives to Antibiotics and Beyond)
Show Figures

Figure 1

Figure 1
<p>Life cycle of a virulent tailed phage (not to scale). (1) The phage collides with the bacterial cell; (2) the phage binds to cell receptors; (3) the phage is irreversibly bound and injects nucleic acid into the cell via the tail tube, where it is transcribed and translated; (4) many progeny phages are produced within intact cells; (5) endolysins degrade the host bacterial cell wall, which loses its structural integrity and ruptures due to the osmotic pressure, releasing the progeny phages.</p>
Full article ">Figure 2
<p>Diagram of the typical cell wall and peptidoglycan structure of bacteria, including the endolysin cleavage sites. The peptidoglycan is composed of repeating sugar units, <span class="html-italic">N</span>-acetylglucosamine (GlcNAc) and <span class="html-italic">N</span>-acetylmuramic acid (MurNAc), which are cross-linked via an interpeptide bridge between the <span class="html-italic">meso</span>-diaminopimelic acid (m-DAP) and <span class="html-small-caps">d</span>-alanine (<span class="html-small-caps">d</span>-Ala) residues of adjacent tetrapeptide chains. The chains also contain <sub>L</sub>-alanine (<span class="html-small-caps">l</span>-Ala) and <sub>D</sub>-glutamic acid (<span class="html-small-caps">d</span>-Glu). Gram-negative bacteria contain an outer membrane (OM) structure not present in Gram-positive bacteria. Both contain an inner membrane (IM) structure. The cleaved bonds and major classifications of endolysin are indicated: (1) <span class="html-italic">N</span>-acetylmuramoyl-<span class="html-small-caps">l</span>-alanine amidase; (2–4) various endopeptidases; (5) N-acetyl-β-<span class="html-small-caps">d</span>-glucosaminidase; (6) N-acetyl-β-<span class="html-small-caps">d</span>-muramidase (lysozyme).</p>
Full article ">
11 pages, 2604 KiB  
Article
Protective Effects of Bacteriophages against Aeromonas hydrophila Causing Motile Aeromonas Septicemia (MAS) in Striped Catfish
by Tuan Son Le, Thi Hien Nguyen, Hong Phuong Vo, Van Cuong Doan, Hong Loc Nguyen, Minh Trung Tran, Trong Tuan Tran, Paul C. Southgate and D. İpek Kurtböke
Antibiotics 2018, 7(1), 16; https://doi.org/10.3390/antibiotics7010016 - 25 Feb 2018
Cited by 69 | Viewed by 9425
Abstract
To determine the effectivity of bacteriophages in controlling the mass mortality of striped catfish (Pangasianodon hypophthalmus) due to infections caused by Aeromonas spp. in Vietnamese fish farms, bacteriophages against pathogenic Aeromonas hydrophila were isolated. A. hydrophila-phage 2 and A. hydrophila [...] Read more.
To determine the effectivity of bacteriophages in controlling the mass mortality of striped catfish (Pangasianodon hypophthalmus) due to infections caused by Aeromonas spp. in Vietnamese fish farms, bacteriophages against pathogenic Aeromonas hydrophila were isolated. A. hydrophila-phage 2 and A. hydrophila-phage 5 were successfully isolated from water samples from the Saigon River of Ho Chi Minh City, Vietnam. These phages, belonging to the Myoviridae family, were found to have broad activity spectra, even against the tested multiple-antibiotic-resistant Aeromonas isolates. The latent periods and burst size of phage 2 were 10 min and 213 PFU per infected host cell, respectively. The bacteriophages proved to be effective in inhibiting the growth of the Aeromonas spp. under laboratory conditions. Phage treatments applied to the pathogenic strains during infestation of catfish resulted in a significant improvement in the survival rates of the tested fishes, with up to 100% survival with MOI 100, compared to 18.3% survival observed in control experiments. These findings illustrate the potential for using phages as an effective bio-treatment method to control Motile Aeromonas Septicemia (MAS) in fish farms. This study provides further evidence towards the use of bacteriophages to effectively control disease in aquaculture operations. Full article
(This article belongs to the Special Issue Bacteriophages: Alternatives to Antibiotics and Beyond)
Show Figures

Figure 1

Figure 1
<p>Plaque formation and microphotograph of <span class="html-italic">A. hydrophila</span> phages. (<b>a</b>,<b>b</b>) Φ2 and (<b>c</b>,<b>d</b>) phage Φ5.</p>
Full article ">Figure 2
<p>Restriction enzyme-digested fragments of the genomic DNA of <span class="html-italic">A. hydrophila</span>-phage 2. Footnote: Lane M: 1kb Plus Opti-DNA Marker (ABM, Canada); Lane L1: genomic DNA of Φ2; Lanes L2–L8: genomic DNA of Φ2 digested with EcoRV; EcoRI; Ncol; SalI; MspI; XmnI; KpnI, restriction enzymes respectively.</p>
Full article ">Figure 3
<p>(<b>a</b>) Adsorption rate and (<b>b</b>) one-step growth curves of Φ2.</p>
Full article ">Figure 3 Cont.
<p>(<b>a</b>) Adsorption rate and (<b>b</b>) one-step growth curves of Φ2.</p>
Full article ">Figure 4
<p>Inactivation of <span class="html-italic">A. hydrophila</span> N17 by the phages (<b>a</b>) Φ2 and (<b>b</b>) Φ5 at different MOI (0.01, 0.1 and 1).</p>
Full article ">Figure 5
<p>Cumulative mortality rates (%) of striped catfishes obtained in challenging experiments using <span class="html-italic">A. hydrophia</span> N17 and the phage cocktail at the different MOIs (0.01, 0.1, and 1). The ratio of Φ2 to Φ5 in a phage cocktail was 1:1.</p>
Full article ">
23 pages, 1810 KiB  
Review
Bacteriophage Interactions with Marine Pathogenic Vibrios: Implications for Phage Therapy
by Panos G. Kalatzis, Daniel Castillo, Pantelis Katharios and Mathias Middelboe
Antibiotics 2018, 7(1), 15; https://doi.org/10.3390/antibiotics7010015 - 24 Feb 2018
Cited by 73 | Viewed by 11425
Abstract
A global distribution in marine, brackish, and freshwater ecosystems, in combination with high abundances and biomass, make vibrios key players in aquatic environments, as well as important pathogens for humans and marine animals. Incidents of Vibrio-associated diseases (vibriosis) in marine aquaculture are [...] Read more.
A global distribution in marine, brackish, and freshwater ecosystems, in combination with high abundances and biomass, make vibrios key players in aquatic environments, as well as important pathogens for humans and marine animals. Incidents of Vibrio-associated diseases (vibriosis) in marine aquaculture are being increasingly reported on a global scale, due to the fast growth of the industry over the past few decades years. The administration of antibiotics has been the most commonly applied therapy used to control vibriosis outbreaks, giving rise to concerns about development and spreading of antibiotic-resistant bacteria in the environment. Hence, the idea of using lytic bacteriophages as therapeutic agents against bacterial diseases has been revived during the last years. Bacteriophage therapy constitutes a promising alternative not only for treatment, but also for prevention of vibriosis in aquaculture. However, several scientific and technological challenges still need further investigation before reliable, reproducible treatments with commercial potential are available for the aquaculture industry. The potential and the challenges of phage-based alternatives to antibiotic treatment of vibriosis are addressed in this review. Full article
(This article belongs to the Special Issue Bacteriophages: Alternatives to Antibiotics and Beyond)
Show Figures

Figure 1

Figure 1
<p>Massive mortalities caused by vibriosis in different developmental stages. (<b>a</b>,<b>b</b>) cultured European seabass, <span class="html-italic">Dicentrarchus labrax</span>, (<b>c</b>) cultured European seabass, <span class="html-italic">Dicentrarchus labrax</span> fry and (<b>d</b>) cultured gilthead sea bream, <span class="html-italic">Sparus aurata</span> larvae in the hatchery.</p>
Full article ">Figure 2
<p>Facilities for live feed production from a commercial fish farm unit. (<b>a</b>) <span class="html-italic">Artemia salina</span> in culture tanks with vigorous aeration, where the native presumptive <span class="html-italic">Vibrio</span> load is regularly estimated between 10<sup>7</sup> and 10<sup>8</sup> cells per mL; (<b>b</b>) <span class="html-italic">Brachionus plicatilis</span> culture tanks, where the native presumptive <span class="html-italic">Vibrio</span> load is regularly between 10<sup>2</sup> and 10<sup>8</sup> cells per mL [<a href="#B103-antibiotics-07-00015" class="html-bibr">103</a>].</p>
Full article ">Figure 3
<p>Overview of the main phage defense mechanisms in bacteria. Prevention of viral attachment on the bacterial surface can be achieved by mutating or masking the receptors, as well as downregulation of receptor expression, orchestrated by quorum sensing (QS). DNA injection may be successfully averted by superinfection exclusion (Sie) mechanisms. If phage DNA enters the bacterial host, its digestion can be catalyzed by R-M mechanism and CRISPR-Cas arrays systems. Deliberate death of the infected cell (abortive infection) constitutes another strategy against viral predators, where prevention of phage proliferation reduces spreading of the infection to the rest of the population.</p>
Full article ">
12 pages, 6193 KiB  
Article
Fragment-Based Discovery of Inhibitors of the Bacterial DnaG-SSB Interaction
by Zorik Chilingaryan, Stephen J. Headey, Allen T. Y. Lo, Zhi-Qiang Xu, Gottfried Otting, Nicholas E. Dixon, Martin J. Scanlon and Aaron J. Oakley
Antibiotics 2018, 7(1), 14; https://doi.org/10.3390/antibiotics7010014 - 22 Feb 2018
Cited by 14 | Viewed by 4767
Abstract
In bacteria, the DnaG primase is responsible for synthesis of short RNA primers used to initiate chain extension by replicative DNA polymerase(s) during chromosomal replication. Among the proteins with which Escherichia coli DnaG interacts is the single-stranded DNA-binding protein, SSB. The C-terminal hexapeptide [...] Read more.
In bacteria, the DnaG primase is responsible for synthesis of short RNA primers used to initiate chain extension by replicative DNA polymerase(s) during chromosomal replication. Among the proteins with which Escherichia coli DnaG interacts is the single-stranded DNA-binding protein, SSB. The C-terminal hexapeptide motif of SSB (DDDIPF; SSB-Ct) is highly conserved and is known to engage in essential interactions with many proteins in nucleic acid metabolism, including primase. Here, fragment-based screening by saturation-transfer difference nuclear magnetic resonance (STD-NMR) and surface plasmon resonance assays identified inhibitors of the primase/SSB-Ct interaction. Hits were shown to bind to the SSB-Ct-binding site using 15N–1H HSQC spectra. STD-NMR was used to demonstrate binding of one hit to other SSB-Ct binding partners, confirming the possibility of simultaneous inhibition of multiple protein/SSB interactions. The fragment molecules represent promising scaffolds on which to build to discover new antibacterial compounds. Full article
(This article belongs to the Special Issue Bacterial DNA Replication and Replication Inhibitors)
Show Figures

Figure 1

Figure 1
<p>Superimposition of <sup>15</sup>N–<sup>1</sup>H HSQC spectra of DnaGC. The protein spectrum in the absence of fragment in black is compared with its spectrum after addition of fragment <b>4</b> (structure shown) in red. Representative assignments of resonances that showed the highest weighted chemical shift perturbation (CSP) (<a href="#app1-antibiotics-07-00014" class="html-app">Figure S3d</a>) are shown.</p>
Full article ">Figure 2
<p>Modeled orientation of fragment <b>4</b>. (<b>a</b>) The docked orientation of fragment <b>4</b> (green carbon atoms) in the single-stranded DNA-binding (SSB)-Ct binding pocket of DnaGC (gray carbon atoms). (<b>b</b>) A schematic representation of interactions between fragment <b>4</b> and its binding pocket. In all structural figures, the protein was visualized using visual molecular dynamics (VMD) [<a href="#B31-antibiotics-07-00014" class="html-bibr">31</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) Structure of hits with binding affinities for further optimization. (<b>b</b>) <sup>15</sup>N–<sup>1</sup>H HSQC titration of fragment <b>4</b>. Binding affinities (<span class="html-italic">K</span><sub>D</sub> values) were derived from the change in chemical shift, Δδ, with incremental additions of ligand.</p>
Full article ">Figure 4
<p>Visualization of binding of compound <b>5</b>. (<b>a</b>) The lowest energy binding poses of <b>5</b> (green carbon atoms) bound to DnaGC (gray carbon atoms). (<b>b</b>) Schematic representation of residues involved in interaction with compound <b>5</b>.</p>
Full article ">Figure 5
<p>1D <sup>19</sup>F-NMR spectra of compound <b>5</b> at 1 mM in the presence (red trace) and absence (blue trace) of 50 μM DnaGC.</p>
Full article ">Figure 6
<p>(<b>a</b>) Saturation transfer difference (STD) spectrum of compound <b>6</b> using DnaG-RCD. In red is a 1D <sup>1</sup>H-NMR reference spectrum, overlaid with a STD spectrum (blue). (<b>b</b>) Overlay of <sup>15</sup>N–<sup>1</sup>H HSQC spectra of <sup>15</sup>N-DnaGC (black) with <b>5</b> (blue) and <b>6</b> (red), each at 1 mM. The apo-protein spectrum is shown in black. Representative assignments of resonances that showed the highest weighted CSP (<a href="#app1-antibiotics-07-00014" class="html-app">Figure S3e,f</a>) are shown.</p>
Full article ">Figure 7
<p>(<b>a</b>) Docked binding pose of <b>6</b> (green carbon atoms) bound to DnaGC (gray carbon atoms). (<b>b</b>) Schematic representation of interactions.</p>
Full article ">Figure 8
<p>Schematic representation of optimization of fragment <b>4</b>. The red labeled groups were added during fragment-to-hit optimization. LE: Ligand efficiency (∆<span class="html-italic">G</span>/[number of heavy atoms]), n.d.: not determined.</p>
Full article ">Figure 9
<p>1D <sup>19</sup>F-NMR spectra. The blue spectrum is of fragment <b>4</b> alone and its spectrum in the presence of <span class="html-italic">E. coli</span> χ is shown in red.</p>
Full article ">
15 pages, 4649 KiB  
Article
Efficacy of an Optimised Bacteriophage Cocktail to Clear Clostridium difficile in a Batch Fermentation Model
by Janet Y. Nale, Tamsin A. Redgwell, Andrew Millard and Martha R. J. Clokie
Antibiotics 2018, 7(1), 13; https://doi.org/10.3390/antibiotics7010013 - 13 Feb 2018
Cited by 69 | Viewed by 9730
Abstract
Clostridium difficile infection (CDI) is a major cause of infectious diarrhea. Conventional antibiotics are not universally effective for all ribotypes, and can trigger dysbiosis, resistance and recurrent infection. Thus, novel therapeutics are needed to replace and/or supplement the current antibiotics. Here, we describe [...] Read more.
Clostridium difficile infection (CDI) is a major cause of infectious diarrhea. Conventional antibiotics are not universally effective for all ribotypes, and can trigger dysbiosis, resistance and recurrent infection. Thus, novel therapeutics are needed to replace and/or supplement the current antibiotics. Here, we describe the activity of an optimised 4-phage cocktail to clear cultures of a clinical ribotype 014/020 strain in fermentation vessels spiked with combined fecal slurries from four healthy volunteers. After 5 h, we observed ~6-log reductions in C. difficile abundance in the prophylaxis regimen and complete C. difficile eradication after 24 h following prophylactic or remedial regimens. Viability assays revealed that commensal enterococci, bifidobacteria, lactobacilli, total anaerobes, and enterobacteria were not affected by either regimens, but a ~2-log increase in the enterobacteria, lactobacilli, and total anaerobe abundance was seen in the phage-only-treated vessel compared to other treatments. The impact of the phage treatments on components of the microbiota was further assayed using metagenomic analysis. Together, our data supports the therapeutic application of our optimised phage cocktail to treat CDI. Also, the increase in specific commensals observed in the phage-treated control could prevent further colonisation of C. difficile, and thus provide protection from infection being able to establish. Full article
(This article belongs to the Special Issue Bacteriophages: Alternatives to Antibiotics and Beyond)
Show Figures

Figure 1

Figure 1
<p>Contributory culturable bacterial counts from each of the individual donors and final cumulative counts of each bacterium added to the fermentation vessels. The bacteria present in the fecal sample of each donor were determined by recovery on selective medium for each bacterial grouping, after which, the samples were mixed together in relatively equal amounts and used to prime the fermentation vessels. The data was analysed using GraphPad Prism 7. Error bars are SEMs of three biological replicates.</p>
Full article ">Figure 2
<p>Impact of phage treatment on <span class="html-italic">C. difficile</span> and other components of the gut microbiota. The impact of phage treatment was ascertained by recovering the bacteria on selective media for (<b>A</b>) <span class="html-italic">C. difficile</span>; (<b>B</b>) bifidobacteria; (<b>C</b>) enterococci; (<b>D</b>) enterobacteria; (<b>E</b>) lactobacilli; (<b>F</b>) total anaerobes<span class="html-italic">.</span> The bacterial counts of the different treatment vessels and time points are presented. Black lines, vessel 1, untreated slurries; red lines, vessel 2, <span class="html-italic">C. difficile</span> control; green lines, vessel 3, phage-only-treated control; blue lines, vessel 4, prophylaxis regimen, and purple lines, vessel 5, remedial regimen. The data was analysed using GraphPad Prism 7. Error bars are SEMs of 3 biological replicates.</p>
Full article ">Figure 3
<p>Analysis of the 10 most abundant taxa from Archea, Bacteria, and Viruses as ascertained by the metagenomics data. Total genomic DNA was extracted at 24 h time point from (<b>A</b>) vessel 1, untreated slurries; (<b>B</b>) vessel 2, <span class="html-italic">C. difficile</span> control; (<b>C</b>) vessel 3, phage-only-treated control; (<b>D</b>) vessel 4, prophylaxis regimen; (<b>E</b>) vessel 5, remedial regimen. The samples were prepared using NexteraXT sample preparation kit and sequenced on MiSeq platform using V2 (2 × 250 bp) chemistry. The resulting fastq files were trimmed with Sickle, and the metagenomes were assembled using megahit. An overview of the 10 most abundant taxa: Phyla (P), Classes (C), order (O), and family (F) are shown for each treatment vessel, as visualised using Pavian. The percent reads mapped to Archaea, Bacteria, and Viruses in the vessels at 24 h are shown in (<b>Fi</b>), (<b>Fii</b>), and (<b>Fiii</b>), respectively.</p>
Full article ">
15 pages, 1970 KiB  
Article
Diversification of Secondary Metabolite Biosynthetic Gene Clusters Coincides with Lineage Divergence in Streptomyces
by Mallory J. Choudoir, Charles Pepe-Ranney and Daniel H. Buckley
Antibiotics 2018, 7(1), 12; https://doi.org/10.3390/antibiotics7010012 - 13 Feb 2018
Cited by 36 | Viewed by 6492
Abstract
We have identified Streptomyces sister-taxa which share a recent common ancestor and nearly identical small subunit (SSU) rRNA gene sequences, but inhabit distinct geographic ranges demarcated by latitude and have sufficient genomic divergence to represent distinct species. Here, we explore the evolutionary dynamics [...] Read more.
We have identified Streptomyces sister-taxa which share a recent common ancestor and nearly identical small subunit (SSU) rRNA gene sequences, but inhabit distinct geographic ranges demarcated by latitude and have sufficient genomic divergence to represent distinct species. Here, we explore the evolutionary dynamics of secondary metabolite biosynthetic gene clusters (SMGCs) following lineage divergence of these sister-taxa. These sister-taxa strains contained 310 distinct SMGCs belonging to 22 different gene cluster classes. While there was broad conservation of these 22 gene cluster classes among the genomes analyzed, each individual genome harbored a different number of gene clusters within each class. A total of nine SMGCs were conserved across nearly all strains, but the majority (57%) of SMGCs were strain-specific. We show that while each individual genome has a unique combination of SMGCs, this diversity displays lineage-level modularity. Overall, the northern-derived (NDR) clade had more SMGCs than the southern-derived (SDR) clade (40.7 ± 3.9 and 33.8 ± 3.9, mean and S.D., respectively). This difference in SMGC content corresponded with differences in the number of predicted open reading frames (ORFs) per genome (7775 ± 196 and 7093 ± 205, mean and S.D., respectively) such that the ratio of SMGC:ORF did not differ between sister-taxa genomes. We show that changes in SMGC diversity between the sister-taxa were driven primarily by gene acquisition and deletion events, and these changes were associated with an overall change in genome size which accompanied lineage divergence. Full article
(This article belongs to the Special Issue Actinomycetes: The Antibiotics Producers)
Show Figures

Figure 1

Figure 1
<p>The northern-derived (NDR) and southern-derived (SDR) clades are closely related sister-taxa and yet were isolated from soils of different latitude. The un-rooted tree was constructed from multiple whole genome alignments with maximum likelihood and a GTRGAMMA model of evolution. Scale bar represents nucleotide substitutions per site. Colored branches depict the northern-derived (NDR) and southern-derived (SDR) clades. Strain names reflect the sample site they were isolated from (<a href="#app1-antibiotics-07-00012" class="html-app">Table S1</a>). Genome NBRC 13350 is the publically available type strain <span class="html-italic">Streptomyces griseus</span> subsp. <span class="html-italic">griseus</span> NBRC 13350. Sample locations are shown in the right panel and labeled with the site code. Circles are colored to reflect the geographic distribution of clades. (Figure modified from [<a href="#B29-antibiotics-07-00012" class="html-bibr">29</a>]).</p>
Full article ">Figure 2
<p>NDR strains have more secondary metabolite biosynthetic gene clusters (SMGCs) than SDR strains (<span class="html-italic">t</span>-test, <span class="html-italic">p</span> &lt; 0.001). (<b>a</b>). Bars indicate the number of SMGCs identified in each genome and are colored according to clade affiliation, and genome names reflect the site of isolation as identified in <a href="#app1-antibiotics-07-00012" class="html-app">Table S1</a>; (<b>b</b>). Kernal density plot shows the distribution of SMGC length (bp).</p>
Full article ">Figure 3
<p>A total of 22 SMGC classes were observed in NDR and SDR genomes by antiSMASH [<a href="#B32-antibiotics-07-00012" class="html-bibr">32</a>]. The tree reflects phylogenetic relationships between <span class="html-italic">Streptomyces</span> sister-taxa genomes and was constructed from multiple whole genome alignments (see <a href="#antibiotics-07-00012-f001" class="html-fig">Figure 1</a>). Scale bar represents nucleotide substitutions per site. Tree branches are colored according to clade affiliation. Bars depict the number of gene clusters belonging to each class for each genome. Colors illustrate gene cluster class as provided by the legend. Asterisks note gene cluster classes that are significantly enriched between clades (<span class="html-italic">t</span>-test and Bonferonni correction for multiple comparisons, <span class="html-italic">p</span> &lt; 0.002).</p>
Full article ">Figure 4
<p>The frequency distribution of SMGCs across strains shows that most SMGCs are strain-specific and fewer are species-specific. Results are shown both for NDR and SDR. (<b>a</b>) and for all 24 genomes; (<b>b</b>). Non-redundant orthologous SMGCs were defined using our annotation-independent approach (see Materials and Methods).</p>
Full article ">Figure 5
<p>We identified 310 non-redundant distinct SMGCs using our annotation-independent gene clustering approach (see Materials and Methods). Each point represents a unique SMGC from a single genome, and colors correspond to clade affiliation. SMGCs with a similar gene composition are clustered spatially, and cluster membership is depicted with polygons. The same data is presented in a different network diagram in <a href="#app1-antibiotics-07-00012" class="html-app">Figure S1</a>.</p>
Full article ">Figure 6
<p>A total of nine core (i.e., conserved in ≥80% of genomes) SMGCs were found in both NDR and SDR. The NDR clade had 11 core SMGCs and the SDR clade had 15 core SMGCs. The tree reflects phylogenetic relationships between <span class="html-italic">Streptomyces</span> sister-taxa genomes and was constructed from multiple whole genome alignments (see <a href="#antibiotics-07-00012-f001" class="html-fig">Figure 1</a>). Scale bar represents nucleotide substitutions per site. Tree branches are colored according to clade affiliation. Core orthologous SMGCs (depicted by colored circles) were determined using the antiSMASH [<a href="#B32-antibiotics-07-00012" class="html-bibr">32</a>] MIBiG annotation pipeline or were defined using our annotation-independent approach (see Materials and Method). Colors correspond to SMGC class (see legend), and natural product annotations are labeled if available.</p>
Full article ">Figure 7
<p>Gene content of core SMGCs vary within and between clades as a result of gene acquisition and deletion events. Panels depict the gene content (i.e., genetic architecture) of core SMGCs (i.e., conserved in ≥80% of genomes), the NDR-specific SMGC core, and the SDR-specific SMGC core. Black bars within the panels represent orthologous genes. The tree reflects phylogenetic relationships between <span class="html-italic">Streptomyces</span> sister-taxa genomes and was constructed from multiple whole genome alignments (see <a href="#antibiotics-07-00012-f001" class="html-fig">Figure 1</a>). Scale bar represents nucleotide substitutions per site. Panel colors correspond to SMGC class (see legend). Panels are labeled with the SMGC cluster membership (see <a href="#app1-antibiotics-07-00012" class="html-app">Table S3</a>) defined using our annotation-independent approach (see Materials and Methods).</p>
Full article ">
7 pages, 410 KiB  
Article
Survey of Nonprescription Medication and Antibiotic Use in Patients with Stevens-Johnson Syndrome, Toxic Epidermal Necrolysis, and Overlap Syndrome
by Katherine J. Sullivan, Meghan N. Jeffres, Robert P. Dellavalle, Robert Valuck and Heather D. Anderson
Antibiotics 2018, 7(1), 11; https://doi.org/10.3390/antibiotics7010011 - 1 Feb 2018
Cited by 3 | Viewed by 6744
Abstract
Stevens-Johnson syndrome (SJS), toxic epidermal necrolysis (TEN), and overlap syndrome (SJS-TEN) are rare, serious skin and mucosa break-down conditions frequently associated with antibiotic use. The role of nonprescription medications alone, or in combination with antibiotics in triggering SJS/TEN, is largely unknown. This study [...] Read more.
Stevens-Johnson syndrome (SJS), toxic epidermal necrolysis (TEN), and overlap syndrome (SJS-TEN) are rare, serious skin and mucosa break-down conditions frequently associated with antibiotic use. The role of nonprescription medications alone, or in combination with antibiotics in triggering SJS/TEN, is largely unknown. This study summarized data collected from patient surveys about nonprescription and antibiotic use prior to a SJS/TEN diagnosis. The survey was administered online to members of the U.S. SJS Foundation who had been diagnosed with SJS/TEN or were the parent of a child who had been diagnosed with SJS/TEN. Respondents were asked about nonprescription medications taken within the year before diagnosis, and the approximate point in time before diagnosis that they had taken them. They were also asked about specific prescription medications, including antibiotics, that they took before diagnosis. An estimated 4500 patients received an invitation to complete the survey. 251 patients completed it, resulting in a response rate of 5.6%. The mean age of respondents was 43 years (SD (standard deviation) = 17.3) and 70% were female. 32.3% of respondents indicated that a prescription antibiotic triggered their reaction. 14.1% indicated a nonprescription medication had triggered their SJS/TEN, and 18.1% said a nonprescription medication may have triggered their SJS/TEN. 85.5% of respondents said they took a nonprescription medication within three months of their SJS/TEN diagnosis. Of those respondents who reported that an antibiotic triggered their SJS/TEN, 35.2% reported taking a nonprescription medication within the three months prior to their diagnosis. This survey captured valuable information about nonprescription and antibiotic use in SJS/TEN patients. It is important for future studies to estimate the impact of antibiotics on SJS/TEN, and account for nonprescription medication use in that relationship. Full article
(This article belongs to the Special Issue Top 35 of Antibiotics Travel Awards 2017)
Show Figures

Figure 1

Figure 1
<p>Antibiotics most often reported by respondents as the trigger of their SJS/TEN (<span class="html-italic">n</span> = 81); SMZ-TMP = sulfamethoxazole-trimethoprim.</p>
Full article ">Figure 2
<p>Nonprescription medications reported by respondents as possible triggers of their SJS/TEN (<span class="html-italic">n</span> = 53).</p>
Full article ">
8 pages, 225 KiB  
Article
Assessing the Knowledge, Attitudes and Behaviors of Human and Animal Health Students towards Antibiotic Use and Resistance: A Pilot Cross-Sectional Study in the UK
by Oliver James Dyar, Holly Hills, Lara-Turiya Seitz, Alex Perry and Diane Ashiru-Oredope
Antibiotics 2018, 7(1), 10; https://doi.org/10.3390/antibiotics7010010 - 30 Jan 2018
Cited by 60 | Viewed by 10713
Abstract
The Global Action Plan on Antimicrobial Resistance highlights the importance of training all healthcare professionals. No study has assessed patterns of students’ knowledge, attitudes and practices concerning antibiotic use simultaneously across different healthcare course types. We conducted a cross-sectional multi-center survey among UK [...] Read more.
The Global Action Plan on Antimicrobial Resistance highlights the importance of training all healthcare professionals. No study has assessed patterns of students’ knowledge, attitudes and practices concerning antibiotic use simultaneously across different healthcare course types. We conducted a cross-sectional multi-center survey among UK students. The survey was advertised through local survey coordinators at 25 universities. The online survey was accessible from 10th October to 17th November 2016 (before European Antibiotic Awareness Day). A total of 255 students from 25 universities participated, including students on medicine, pharmacy, nursing, physician associate, dentistry and veterinary medicine courses. Antibiotic resistance was considered to be a more important global challenge than climate change, obesity or food security (p < 0.001). Most students (95%) believed that antibiotic resistance will be a problem for their future practice, but fewer (69%) thought that the antibiotics they will prescribe, administer or dispense will contribute to the problem. A fifth of students felt they had sufficient knowledge of antibiotic use for their future work. Our exploratory study suggests that UK human and animal healthcare students are aware of the importance of antibiotic resistance, but many still have certain misconceptions. Campaigns and improved educational efforts applying behavioral insights methodology could address these. Full article
(This article belongs to the Special Issue Top 35 of Antibiotics Travel Awards 2017)
16 pages, 1965 KiB  
Article
Use of a Regression Model to Study Host-Genomic Determinants of Phage Susceptibility in MRSA
by Henrike Zschach, Mette V. Larsen, Henrik Hasman, Henrik Westh, Morten Nielsen, Ryszard Międzybrodzki, Ewa Jończyk-Matysiak, Beata Weber-Dąbrowska and Andrzej Górski
Antibiotics 2018, 7(1), 9; https://doi.org/10.3390/antibiotics7010009 - 29 Jan 2018
Cited by 5 | Viewed by 4627
Abstract
Staphylococcus aureus is a major agent of nosocomial infections. Especially in methicillin-resistant strains, conventional treatment options are limited and expensive, which has fueled a growing interest in phage therapy approaches. We have tested the susceptibility of 207 clinical S. aureus strains to 12 [...] Read more.
Staphylococcus aureus is a major agent of nosocomial infections. Especially in methicillin-resistant strains, conventional treatment options are limited and expensive, which has fueled a growing interest in phage therapy approaches. We have tested the susceptibility of 207 clinical S. aureus strains to 12 (nine monovalent) different therapeutic phage preparations and subsequently employed linear regression models to estimate the influence of individual host gene families on resistance to phages. Specifically, we used a two-step regression model setup with a preselection step based on gene family enrichment. We show that our models are robust and capture the data’s underlying signal by comparing their performance to that of models build on randomized data. In doing so, we have identified 167 gene families that govern phage resistance in our strain set and performed functional analysis on them. This revealed genes of possible prophage or mobile genetic element origin, along with genes involved in restriction-modification and transcription regulators, though the majority were genes of unknown function. This study is a step in the direction of understanding the intricate host-phage relationship in this important pathogen with the outlook to targeted phage therapy applications. Full article
(This article belongs to the Special Issue Bacteriophages: Alternatives to Antibiotics and Beyond)
Show Figures

Figure 1

Figure 1
<p>All-against-all matrix of the genetic distance between the 207 methicillin-resistant <span class="html-italic">Staphylococcus aureus</span> (MRSA) strains used for this study. Distance is calculated as 1-orthoANI and represented as color, where blue corresponds to lower and red corresponds to greater distance. The assignment of strains to partitions is marked on the right margin.</p>
Full article ">Figure 2
<p>Abundance of gene families in the 207 strains. The peak depicted in the histogram is slightly higher than the number of housekeeping genes, 1.777, since the bin is wider than 1.</p>
Full article ">Figure 3
<p>Stacked histogram of <span class="html-italic">p</span>-value distributions across the five folds for the interaction with phage P4/6409. The density is shown instead of counts to account for fold 1 having a 100 times less <span class="html-italic">p</span>-values compared to the other folds, since it does not include partition 1 and therefore did not need to be subsampled. <b>Left</b>: Real data. <b>Right</b>: Permuted data.</p>
Full article ">Figure 4
<p>Heat map of the regression weights for the final model of phage P4/6409. Columns are gene families, rows are cross validation folds. The color indicates the value and direction of each weight, with blue being strongly positive and red being strongly negative. Weights with low values are white. Results were comparable for other phages with the exception of 1N/80, A3/R, and mix MS-1 (see <a href="#antibiotics-07-00009-t002" class="html-table">Table 2</a>).</p>
Full article ">Figure 5
<p>Functional annotation categories of the eggNOGs matching to the set of significant genes across all nine phages.</p>
Full article ">Figure 6
<p>Histogram depicting the number of phage models where a given gene family was identified significant.</p>
Full article ">
25 pages, 1303 KiB  
Article
Phage-Bacterial Dynamics with Spatial Structure: Self Organization around Phage Sinks Can Promote Increased Cell Densities
by James J. Bull, Kelly A. Christensen, Carly Scott, Benjamin R. Jack, Cameron J. Crandall and Stephen M. Krone
Antibiotics 2018, 7(1), 8; https://doi.org/10.3390/antibiotics7010008 - 29 Jan 2018
Cited by 35 | Viewed by 5518
Abstract
Bacteria growing on surfaces appear to be profoundly more resistant to control by lytic bacteriophages than do the same cells grown in liquid. Here, we use simulation models to investigate whether spatial structure per se can account for this increased cell density in [...] Read more.
Bacteria growing on surfaces appear to be profoundly more resistant to control by lytic bacteriophages than do the same cells grown in liquid. Here, we use simulation models to investigate whether spatial structure per se can account for this increased cell density in the presence of phages. A measure is derived for comparing cell densities between growth in spatially structured environments versus well mixed environments (known as mass action). Maintenance of sensitive cells requires some form of phage death; we invoke death mechanisms that are spatially fixed, as if produced by cells. Spatially structured phage death provides cells with a means of protection that can boost cell densities an order of magnitude above that attained under mass action, although the effect is sometimes in the opposite direction. Phage and bacteria self organize into separate refuges, and spatial structure operates so that the phage progeny from a single burst do not have independent fates (as they do with mass action). Phage incur a high loss when invading protected areas that have high cell densities, resulting in greater protection for the cells. By the same metric, mass action dynamics either show no sustained bacterial elevation or oscillate between states of low and high cell densities and an elevated average. The elevated cell densities observed in models with spatial structure do not approach the empirically observed increased density of cells in structured environments with phages (which can be many orders of magnitude), so the empirical phenomenon likely requires additional mechanisms than those analyzed here. Full article
(This article belongs to the Special Issue Bacteriophages: Alternatives to Antibiotics and Beyond)
Show Figures

Figure 1

Figure 1
<p>The density of cells maintained in the presence of phage is often increased by spatial structure. Shown in each panel are the <math display="inline"> <semantics> <msub> <mi>A</mi> <mi>g</mi> </msub> </semantics> </math> values, giving the fold increase in cell density over that with mass action. <math display="inline"> <semantics> <msub> <mi>A</mi> <mi>g</mi> </msub> </semantics> </math> values are greatly influenced by EPS levels and burst sizes, exceeding 10 only in the upper right quadrant, with large bursts and high EPS densities, and then only for some values of diffusion and cell reproduction rate. Values within each panel give average <math display="inline"> <semantics> <msub> <mi>A</mi> <mi>g</mi> </msub> </semantics> </math> values from 15 trials each using the same burst and EPS levels, with rate of cell reproduction and phage diffusion rate given at the top of each panel; trials leading to extinction of phage or cells are not included in the averages. EPS was assigned randomly to each patch at the start and remained in the patch for the life of the run; superinfection of infected cells was not allowed (<math display="inline"> <semantics> <mrow> <msub> <mi>k</mi> <mi>I</mi> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics> </math>), nor was debris attachment (<math display="inline"> <semantics> <mrow> <msub> <mi>k</mi> <mi>D</mi> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics> </math>). Each trial ran 10,000 time steps, and <span class="html-italic">A</span> was averaged over the last 3000 steps; values are rounded to the nearest integer (values rounded to 1 were often less than 1). A black subscript denotes the number of trials with bacterial and/or phage extinction; a dot indicates that all 15 trials led to extinction. The ‘cell=’ value given above each panel is the probability that an uninfected cell reproduced at each time step; the ‘diffuse=’ value is the fraction of phage that left the patch in each time step. In all trials, the adsorption probability to uninfected cells was <math display="inline"> <semantics> <mrow> <msub> <mi>k</mi> <mi>C</mi> </msub> <mo>=</mo> <mn>0.25</mn> </mrow> </semantics> </math>, and that to EPS was <math display="inline"> <semantics> <mrow> <msub> <mi>k</mi> <mi>E</mi> </msub> <mo>=</mo> <mn>0.35</mn> </mrow> </semantics> </math>.</p>
Full article ">Figure 2
<p>Illustration of self-organization of phages and cells with clumped EPS. White background indicates EPS, yellow is absence of EPS. A green (orange) circle is an uninfected (infected) cell. A blue or black legged icon is a phage (blue is free, black is attached to a cell). Phages are mostly confined to the EPS-free zone and the first row of EPS. Figure was generated from a NetLogo trial with a grid size of 21 × 21, a burst of 20, diffusion step size of 0.45 and attachment probabilities as in <a href="#antibiotics-07-00008-f001" class="html-fig">Figure 1</a>C. There were 32 phage (partially obscured by cells) in the first three rows of EPS; <math display="inline"> <semantics> <mrow> <mi>α</mi> <mi>b</mi> </mrow> </semantics> </math> for the entire grid was 7.26.</p>
Full article ">
2 pages, 208 KiB  
Editorial
Acknowledgement to Reviewers of Antibiotics in 2017
by Antibiotics Editorial Office
Antibiotics 2018, 7(1), 7; https://doi.org/10.3390/antibiotics7010007 - 25 Jan 2018
Viewed by 2673
Abstract
Peer review is an essential part in the publication process, ensuring that Antibiotics maintains high quality standards for its published papers.[...] Full article
6 pages, 179 KiB  
Article
Characteristics of Pediatric Antimicrobial Stewardship Programs: Current Status of the Sharing Antimicrobial Reports for Pediatric Stewardship (SHARPS) Collaborative
by Christopher McPherson, Brian R. Lee, Cindy Terrill, Adam L. Hersh, Jeffrey S. Gerber, Matthew P. Kronman and Jason G. Newland
Antibiotics 2018, 7(1), 4; https://doi.org/10.3390/antibiotics7010004 - 25 Jan 2018
Cited by 22 | Viewed by 5498
Abstract
In response to the growing epidemic of antibiotic-resistant bacterial infections, antimicrobial stewardship programs (ASP) have been rapidly implemented in the United States (US). This study examines the prevalence of the Centers for Disease Control and Prevention’s (CDC) seven core elements of a successful [...] Read more.
In response to the growing epidemic of antibiotic-resistant bacterial infections, antimicrobial stewardship programs (ASP) have been rapidly implemented in the United States (US). This study examines the prevalence of the Centers for Disease Control and Prevention’s (CDC) seven core elements of a successful ASP within a large subset of US Children’s Hospitals. In 2016, a survey was conducted of 52 pediatric hospitals assessing the presence of the seven core elements: leadership commitment, accountability, drug expertise, action, tracking, reporting, and education. Forty-nine hospitals (94%) had established ASPs and 41 hospitals (79%) included all seven core elements. Physician accountability (87%) and a dedicated ASP pharmacist or drug expert (88%) were present in the vast majority of hospitals. However, substantial variability existed in the financial support allotted to these positions. This variability did not predict program actions, tracking, reporting, and education. When compared with previous surveys, these results document a dramatic increase in the prevalence and resources of pediatric stewardship programs, although continued expansion is warranted. Further research is required to understand the feasibility of various core stewardship activities and the impact on patient outcomes in the setting of finite resources. Full article
(This article belongs to the Special Issue Surveillance of Antimicrobial Use and Resistance in Children)
15 pages, 730 KiB  
Article
Reevaluation of the Acute Cystitis Symptom Score, a Self-Reporting Questionnaire. Part I. Development, Diagnosis and Differential Diagnosis
by Jakhongir F. Alidjanov, Kurt G. Naber, Ulugbek A. Abdufattaev, Adrian Pilatz and Florian M. E. Wagenlehner
Antibiotics 2018, 7(1), 6; https://doi.org/10.3390/antibiotics7010006 - 15 Jan 2018
Cited by 20 | Viewed by 5382
Abstract
This study aimed to reevaluate the Acute Cystitis Symptom Score (ACSS). The ACSS is a simple and standardized self-reporting questionnaire for the diagnosis of acute uncomplicated cystitis (AC) assessing typical and differential symptoms, quality of life, and possible changes after therapy in female [...] Read more.
This study aimed to reevaluate the Acute Cystitis Symptom Score (ACSS). The ACSS is a simple and standardized self-reporting questionnaire for the diagnosis of acute uncomplicated cystitis (AC) assessing typical and differential symptoms, quality of life, and possible changes after therapy in female patients with AC. This paper includes literature research, development and evaluation of the ACSS, an 18-item self-reporting questionnaire including (a) six questions about “typical” symptoms of AC, (b) four questions regarding differential diagnoses, (c) three questions on quality of life, and (d) five questions on additional conditions that may affect therapy. The ACSS was evaluated in 228 women (mean age 31.49 ± 11.71 years) in the Russian and Uzbek languages. Measurements of reliability, validity, predictive ability, and responsiveness were performed. Cronbach’s alpha for ACSS was 0.89, split-half reliability was 0.76 and 0.79 for first and second halves, and the correlation between them was 0.87. Mann-Whitney U test revealed a significant difference in scores of the “typical” symptoms between patients and controls (10.50 vs. 2.07, p < 0.001). The optimal threshold score was 6 points, with a 94% sensitivity and 90% specificity to predict AC. The “typical” symptom score decreased significantly when comparing before and after therapy (10.4 and 2.5, p < 0.001). The reevaluated Russian and Uzbek ACSS are accurate enough and can be recommended for clinical studies and practice for initial diagnosis and monitoring the process of the treatment of AC in women. Evaluation in German, UK English, and Hungarian languages was also performed and in other languages evaluation of the ACSS is in progress Full article
Show Figures

Figure 1

Figure 1
<p>Receiver-Operating-Characteristic (ROC) curve analysis of the “Typical” domain of the Acute Cystitis Symptom Score (ACSS) at the first visit with results from 228 subjects (107 patients with acute uncomplicated cystitis (AC), 121 controls without AC).</p>
Full article ">Figure 2
<p>Distribution of 228 subjects (107 patients with AC, 121 controls without AC) according to a summary score of the “Typical” domain of the ACSS at the first visit.</p>
Full article ">
14 pages, 3590 KiB  
Article
Screening of E. coli β-clamp Inhibitors Revealed that Few Inhibit Helicobacter pylori More Effectively: Structural and Functional Characterization
by Preeti Pandey, Vijay Verma, Suman Kumar Dhar and Samudrala Gourinath
Antibiotics 2018, 7(1), 5; https://doi.org/10.3390/antibiotics7010005 - 11 Jan 2018
Cited by 3 | Viewed by 4832
Abstract
The characteristic of interaction with various enzymes and processivity-promoting nature during DNA replication makes β-clamp an important drug target. Helicobacter pylori (H. pylori) have several unique features in DNA replication machinery that makes it different from other microorganisms. To find out [...] Read more.
The characteristic of interaction with various enzymes and processivity-promoting nature during DNA replication makes β-clamp an important drug target. Helicobacter pylori (H. pylori) have several unique features in DNA replication machinery that makes it different from other microorganisms. To find out whether difference in DNA replication proteins behavior accounts for any difference in drug response when compared to E. coli, in the present study, we have tested E. coli β-clamp inhibitor molecules against H. pylori β-clamp. Various approaches were used to test the binding of inhibitors to H. pylori β-clamp including docking, surface competition assay, complex structure determination, as well as antimicrobial assay. Out of five shortlisted inhibitor molecules on the basis of docking score, three molecules, 5-chloroisatin, carprofen, and 3,4-difluorobenzamide were co-crystallized with H. pylori β-clamp and the structures show that they bind at the protein-protein interaction site as expected. In vivo studies showed only two molecules, 5-chloroisatin, and 3,4-difluorobenzamide inhibited the growth of the pylori with MIC values in micro molar range, which is better than the inhibitory effect of the same drugs on E. coli. Therefore, the evaluation of such drugs against H. pylori may explore the possibility to use to generate species-specific pharmacophore for development of new drugs against H. pylori. Full article
(This article belongs to the Special Issue Bacterial DNA Replication and Replication Inhibitors)
Show Figures

Figure 1

Figure 1
<p>SPR sensorgram. SPR sensorgram showing surface competition assay between HpDNA ligase and the small molecules. A qualitative analysis of the in vitro competition between DNA ligase and small molecules (present in solution) for binding to Hpβ-clamp (immobilized on the chip surface) was carried out. For the small molecules (<b>A</b>) 5-chloroisatin (C1); (<b>B</b>) 6-nitroindazole (C2) and (<b>C</b>) (S)-carprofen (C3), a mass of ~6 ng of Hpβ-clamp gets immobilized on the chip surface while for (<b>D</b>) 5-nitroindole (C4) and (<b>E</b>) 3,4-difluorobenzamide (C5), a mass of ~4 ng of Hpβ-clamp gets immobilized on the chip surface. The concentration of ligase was kept the same throughout each experiment with a small molecule. As the concentration of the small molecule was increased, the SPR response decreased.</p>
Full article ">Figure 2
<p>5-chloroisatin interaction with Hpβ-clamp. (<b>A</b>) 2Fo-Fc map, contoured at 1σ, of 5-chloroisatin bound to Hpβ-clamp (PDB ID: 5G4Q). The alignment of the structure of the complex (green) with that of the native (orange) did not yield significant differences in the orientation of interacting residues except for I248; (<b>B</b>) Ligplot of the Hpβ-clamp structure near the bound 5-chloroisatin, showing the predominantly hydrophobic interactions between the protein and the inhibitor. T173 of Hpβ-clamp did form an H-bond with the inhibitor molecule; (<b>C</b>) Structural alignment of Hpβ-clamp (green) and Ecβ-clamp (cyan) complex with ligand 5-chloroisatin. T172 of Ecβ-clamp makes H-bond with ligand while T175 of Hpβ-clamp makes H-bond with the ligand (PDB ID: 4N95); (<b>D</b>) Ligplot of Ecβ-clamp complex with 5-chloroisatin showing the types of interactions between them.</p>
Full article ">Figure 3
<p>(S)-carprofen interaction with β-clamp. (<b>A</b>) 2Fo-Fc map, contoured at 1σ, of (S)-carprofen bound to Hpβ-clamp (PDB ID: 5FXT). The alignment of the structure of the co-crystal of Hpβ-clamp (green) and the inhibitor (olive) with that of the native Hpβ-clamp (orange) showed almost same orientation of interacting molecules in both structures; (<b>B</b>) Ligplot of the Hpβ-clamp structure near the bound (S)-carprofen, showing the hydrophobic interactions between the protein and the inhibitor; (<b>C</b>) Structural alignment of <span class="html-italic">H. pylori</span> (green) and <span class="html-italic">E. coli</span> β-clamp (cyan) complex with ligand (S)-carprofen. T154 of Ecβ-clamp makes H-bond with the ligand apart from other hydrophobic interactions while Hpβ-clamp and ligand interactions are dominated by hydrophobic interactions; and, (<b>D</b>) Ligplot of Ecβ-clamp with bound ligand showing its various interactions with ligand (PDB ID: 4MJR).</p>
Full article ">Figure 4
<p>3, 4-difluorobenzamide interaction with β-clamp. (<b>A</b>) 2Fo-Fc map, contoured at 1σ, of 3, 4-difluorobenzamide in complex with Hpβ-clamp. The alignment of the structure of the co-crystal of Hpβ-clamp (green) and the inhibitor (olive) with that of the native Hpβ-clamp (orange) showed differences in the orientations of residues T175, M370, K176 and I248 between the co-crystal and native structures; (<b>B</b>) Ligplot of the Hpβ-clamp structure near the bound 3, 4-difluorobenzamide, showing the hydrophobic interactions between the protein and the inhibitor; (<b>C</b>) Structural superimposition of Hpβ-clamp (green) and Ecβ-clamp (cyan) complex with ligand 3, 4-difluorobenzamide. In both the cases, the contacts are dominated by hydrophobic interactions however the orientation of ligand molecule is differentiated by a rotation of 180 degree; and, (<b>D</b>) Ligplot of Ecβ-clamp bound to ligand showing the hydrophobic interactions nearby (PDB ID: 4N94).</p>
Full article ">Figure 5
<p>Structure-based sequence alignment. β-clamps of <span class="html-italic">H. pylori</span> and <span class="html-italic">E. coli</span> were compared using structure based sequence alignment. The ligand-interacting residues are highlighted (in green box). In each block, the first line shows conservation indices for positions with a conservation index above 5. The secondary structure prediction is shown in color red (alpha-helix) and blue (beta-strand). The last two lines show consensus amino acid sequence (consensus_aa) and consensus predicted secondary structures (consensus_ss). Consensus predicted secondary structure symbols: alpha-helix:h; beta-strand:e. Consensus amino acid symbols are: conserved amino acid are in uppercase and bold letter; aliphatic (I,V, L): l; aromatic (YHWF); hydrophobic (W,F,Y,M,L,I,V,A,C,T,H): h; alcohol (S,T):o; polar residues (D,E,H,K,N,Q,R,S,T): p; tiny (A,G,C,S):t; small (A,G,C,S,V,N,D,T,P):s; bulky residues (E,F,I,K,L,M,Q,R,W,Y):b; positively charged (K,R,H):+; negatively charged (D,E): -; charged (D,E,K,R,H):c.</p>
Full article ">Figure 6
<p>Antimicrobial activities of different drugs against <span class="html-italic">H. pylori</span>. (<b>A</b>) Anti-<span class="html-italic">H. pylori</span> activities of different drugs (<b>A</b>) 5-chloroisatin (C1) and (<b>B</b>) 3,4-difluorobenzamide (C5) determined by applying the disk diffusion method. Petri dish with drugs containing discs showed inhibition zones for bacterial growth; (<b>C</b>) Anti-<span class="html-italic">H. pylori</span> activities of drugs determined as minimum inhibitory concentrations (MICs) were obtained via the dilution method. The MIC of drug C1 and C5 are 18 µM and 824 µM, respectively. The experiments were performed in triplicates. Error bars show standard deviation of the mean (Mean ± SD).</p>
Full article ">
3523 KiB  
Article
Establishing a System for Testing Replication Inhibition of the Vibrio cholerae Secondary Chromosome in Escherichia coli
by Nadine Schallopp, Sarah Milbredt, Theodor Sperlea, Franziska S. Kemter, Matthias Bruhn, Daniel Schindler and Torsten Waldminghaus
Antibiotics 2018, 7(1), 3; https://doi.org/10.3390/antibiotics7010003 - 23 Dec 2017
Cited by 9 | Viewed by 5835
Abstract
Regulators of DNA replication in bacteria are an attractive target for new antibiotics, as not only is replication essential for cell viability, but its underlying mechanisms also differ from those operating in eukaryotes. The genetic information of most bacteria is encoded on a [...] Read more.
Regulators of DNA replication in bacteria are an attractive target for new antibiotics, as not only is replication essential for cell viability, but its underlying mechanisms also differ from those operating in eukaryotes. The genetic information of most bacteria is encoded on a single chromosome, but about 10% of species carry a split genome spanning multiple chromosomes. The best studied bacterium in this context is the human pathogen Vibrio cholerae, with a primary chromosome (Chr1) of 3 M bps, and a secondary one (Chr2) of about 1 M bps. Replication of Chr2 is under control of a unique mechanism, presenting a potential target in the development of V. cholerae-specific antibiotics. A common challenge in such endeavors is whether the effects of candidate chemicals can be focused on specific mechanisms, such as DNA replication. To test the specificity of antimicrobial substances independent of other features of the V. cholerae cell for the replication mechanism of the V. cholerae secondary chromosome, we establish the replication machinery in the heterologous E. coli system. We characterize an E. coli strain in which chromosomal replication is driven by the replication origin of V. cholerae Chr2. Surprisingly, the E. coli ori2 strain was not inhibited by vibrepin, previously found to inhibit ori2-based replication. Full article
(This article belongs to the Special Issue Bacterial DNA Replication and Replication Inhibitors)
Show Figures

Figure 1

Figure 1
<p>DNA replication in <span class="html-italic">E. coli</span> strains with <span class="html-italic">oriC</span> or <span class="html-italic">ori2</span> driven replication. (<b>A</b>) Flow cytometry analyses of DNA content in rifampicin/cephalexin treated <span class="html-italic">E. coli</span> cells with DNA replication starting at <span class="html-italic">oriC</span> (strain #1; top panel) or <span class="html-italic">V. cholerae ori2</span> (strain #16; bottom panel) [<a href="#B25-antibiotics-07-00003" class="html-bibr">25</a>]. (<b>B</b>) Fluorescence microscopy of <span class="html-italic">E. coli</span> cells harboring a plasmid encoding a TetR-mVenus fusion (pMA289), a FROS array insertion and either <span class="html-italic">oriC</span> (top panel; strain SM112) or <span class="html-italic">ori2</span> (bottom panel; strain SM113). The scale bar is 2 µm. (<b>C</b>) Quantification of fluorescence foci per cell for microscopy shown in B (n = 700). (<b>D</b>) Profile of genome-wide copy numbers based on comparative genomic hybridization (CGH). Grey dots represent values of single probes for the <span class="html-italic">ori2</span>-based strain (#16) with a Loess regression (red line, F = 0.2). For comparison, the Loess regression of the <span class="html-italic">oriC</span>-based strain #1 is shown based on published data [<a href="#B29-antibiotics-07-00003" class="html-bibr">29</a>] (blue line, F = 0.2). Positions of <span class="html-italic">oriC</span> and <span class="html-italic">ori2</span> are indicated and the genomic position as distance from <span class="html-italic">oriC</span>.</p>
Full article ">Figure 2
<p>Transduction efficiency of gene knockouts into <span class="html-italic">oriC</span> and <span class="html-italic">ori2 E. coli</span> strains. Colonies growing on plates supplemented with kanamycin were counted after P1 transduction of KanR-cassettes inserted in <span class="html-italic">dam</span>, <span class="html-italic">hapA</span> or <span class="html-italic">seqA</span> as indicated into <span class="html-italic">E. coli</span> strains, with chromosome replication based on <span class="html-italic">oriC</span> (MG1655 (#1)) or <span class="html-italic">ori2</span> (NZ90). Mean values of three replicates are shown in (<b>A</b>) with the actual numbers and standard deviation given in the table (<b>B</b>).</p>
Full article ">Figure 3
<p>Copy numbers of secondary replicons in <span class="html-italic">E. coli</span> dependent on <span class="html-italic">crtS</span>. (<b>A</b>) Copy number of an <span class="html-italic">ori2</span>-based minichromosome in an <span class="html-italic">E. coli</span> strain without <span class="html-italic">crtS</span> (Strain NZ72) or with <span class="html-italic">crtS</span>-site insertion on the chromosome (Strain NZ140) was measured as inverse of the lag time for cells in medium with high concentrations of ampicillin (500 µg/ mL). Strains carrying an <span class="html-italic">oriF</span>-based replicon and the <span class="html-italic">crtS</span> (Strain NZ141) or no <span class="html-italic">crtS</span> on the chromosome (Strain NZ119) were used as control. Data are the mean of three biological replicates with the indicated standard deviations. (<b>B</b>) Copy number of an <span class="html-italic">ori2</span>-minichromosome (pMA568) analyzed by qPCR-based marker frequency analysis relative to <span class="html-italic">oriC</span>. Mean values of three biological replicates are shown with the respective standard deviations.</p>
Full article ">Figure 4
<p>Conjugation rate of extra replicons to <span class="html-italic">oriC</span> and <span class="html-italic">ori2 E. coli</span> strains. (<b>A</b>) Transfer by conjugation was tested for <span class="html-italic">oriC</span>-minichromosomes carrying a <span class="html-italic">crtS</span> site (pMA200) or no <span class="html-italic">crtS</span> (pMA308) into <span class="html-italic">E. coli</span> with chromosome replication based on <span class="html-italic">oriC</span> (MG1655) or <span class="html-italic">ori2</span> (NZ90). (<b>B</b>) Conjugation of <span class="html-italic">oriF</span>-based replicons with <span class="html-italic">crtS</span> (pMA206) or without (pMA899) into <span class="html-italic">E. coli</span> with <span class="html-italic">oriC</span> (MG1655), <span class="html-italic">ori2</span> (NZ90) or <span class="html-italic">oriC</span> at the native site and <span class="html-italic">ori2</span> at an ectopic site (#15) [<a href="#B25-antibiotics-07-00003" class="html-bibr">25</a>,<a href="#B29-antibiotics-07-00003" class="html-bibr">29</a>]. Mean values of three biological replicates are shown in (<b>A</b>,<b>B</b>) with the actual numbers and standard deviations given in the table (<b>C</b>).</p>
Full article ">Figure 5
<p>Mutation-based analysis of the DnaA box within <span class="html-italic">ori2.</span> (<b>A</b>–<b>C</b>) <span class="html-italic">Ori2</span>-based minichromosomes with different mutations at the DnaA box were tested for their ability to replicate by transformation into <span class="html-italic">E. coli</span> (XL1-Blue) or DH5αλ<span class="html-italic">pir</span>. The latter was used as a control because all replicons carry an oriR6K which allows replication in strains encoding the initiator Pir. Values are ratios between respective colony numbers of three biological replicates with the indicated standard deviations. Relevant sequences are shown in (<b>D</b>) for DnaA boxes wt (pMA87), scrambled (pMA108), inverted (pMA109), deleted (pMA110), a weak DnaA box R3 (pMA111), a medium-strength DnaA box R2 (pMA112), a 5 bp insertion to the left of the DnaA box (pMA115) or the right (pMA114) or a 5 bp deletion to the left (pMA113) or right (pMA116) as indicated. (<b>E</b>) Sequences found by transformation of an assembled <span class="html-italic">ori2</span>-minichromosome with a mix of sequence combinations at positions 2–5 as indicated by “N” in the Primer sequence. DnaA-box sequences from <span class="html-italic">oriC</span> in <span class="html-italic">E. coli</span> are shown for comparison (wt, R2, R3). Six sequences found in the screen are shown with the nucleotides differing from the consensus DnaA box shaded in grey.</p>
Full article ">Figure 6
<p>Effect of vibrepin on <span class="html-italic">ori2</span>-dependent replication. Lag time (time until OD = 0.1 was reached) is shown as the mean value of three replicates, with the respective standard deviations, in medium without vibrepin (blue) or with vibrepin (red). Vibrepin concentrations were 16 μg/mL for <span class="html-italic">E. coli</span> strains (<b>A</b>–<b>B</b>) and 1.6 μg/mL for <span class="html-italic">V. cholerae</span> strains (<b>C</b>). Analyzed <span class="html-italic">E. coli</span> strains replicated their chromosome based on <span class="html-italic">oriC</span> (MG1655) or <span class="html-italic">ori1</span> (NZ90) (<b>A</b>) or carried an extra minireplicon with <span class="html-italic">ori2</span> (pMA100), <span class="html-italic">oriC</span> (pMA106) or <span class="html-italic">oriF</span> (pMA129) [<a href="#B38-antibiotics-07-00003" class="html-bibr">38</a>] (<b>B</b>) as indicated. <span class="html-italic">V. cholerae</span> strains are the standard two-chromosome strain N16961 [<a href="#B6-antibiotics-07-00003" class="html-bibr">6</a>], an engineered derivative of N16961 with fused chromosomes (MCH1) [<a href="#B36-antibiotics-07-00003" class="html-bibr">36</a>] or a natural isolate with fused chromosomes (NSCV1) [<a href="#B37-antibiotics-07-00003" class="html-bibr">37</a>].</p>
Full article ">Figure 7
<p>Construction and characterization of strain NZ138. (<b>A</b>) Scheme of NZ138 construction. Genes are indicated by arrows, origins of replication and the truncated end of <span class="html-italic">mnmG</span> by blocks. Genes of <span class="html-italic">V. cholerae</span> are colored blue, genes of <span class="html-italic">E. coli</span> orange, resistance genes red, FRT-sites black and origins of replication yellow. Green rectangles indicate BsaI and BpiI restriction sites. Binding sites of oligonucleotides are indicated in purple, numbers in brackets are genome positions of 5’-ends of the binding sites. Genomic and linearized DNA is indicated by grey lines, plasmids by circles. (<b>B</b>) Growth, DNA content and protein content of NZ138 compared to MG1655 and NZ90. All strains were grown in LB medium. For growth curves, five replicates for each strain were grown in a 96-well plate at 37 °C. For determination of DNA and protein content, samples were taken in exponential phase and fixed with ethanol. The samples were split, stained with SYTOX Green (DNA) or FITC (protein) and analyzed by flow cytometry.</p>
Full article ">
2019 KiB  
Article
Evidence for Anti-Pseudogymnoascus destructans (Pd) Activity of Propolis
by Soumya Ghosh, Robyn McArthur, Zhi Chao Guo, Rory McKerchar, Kingsley Donkor, Jianping Xu and Naowarat Cheeptham
Antibiotics 2018, 7(1), 2; https://doi.org/10.3390/antibiotics7010002 - 21 Dec 2017
Cited by 6 | Viewed by 6494
Abstract
White-nose syndrome (WNS) in bats, caused by Pseudogymnoascus destructans (Pd), is a cutaneous infection that has devastated North American bat populations since 2007. At present, there is no effective method for controlling this disease. Here, we evaluated the effect of propolis [...] Read more.
White-nose syndrome (WNS) in bats, caused by Pseudogymnoascus destructans (Pd), is a cutaneous infection that has devastated North American bat populations since 2007. At present, there is no effective method for controlling this disease. Here, we evaluated the effect of propolis against Pd in vitro. Using Sabouraud dextrose agar (SDA) medium, approximately 1.7 × 107 conidia spores of the Pd strain M3906-2/mL were spread on each plate and grown to form a consistent lawn. A Kirby–Bauer disk diffusion assay was employed using different concentrations of propolis (1%, 2%, 3%, 4%, 5%, 10%, 15%, 20%, 25%), in plates incubated at 8 °C and 15 °C. At 8 °C and 15 °C, as the concentration of propolis increased, there was an increasing zone of inhibition (ZOI), reaching the highest degree at 10% and 25% concentrations, respectively. A germule suppression assay showed a similar effect on Pd conidia germination. A MALDI-TOF-MS analysis of propolis revealed multiple constituents with a potential anti-Pd activity, including cinnamic acid, p-coumaric acid, and dihydrochalcones, which could be further tested for their individual effects. Our study suggests that propolis or its individual constituents might be suitable products against Pd. Full article
(This article belongs to the Special Issue Top 35 of Antibiotics Travel Awards 2017)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Anti-<span class="html-italic">Pd</span> activity of propolis. (<b>A</b>) Images (i–xi) and (xii–xxii) indicate the activity of propolis at 8 °C and 15 °C, respectively. The black arrowheads indicate the zone of inhibition of <span class="html-italic">Pd</span> when treated with different concentrations of propolis in comparison to water and anhydrous ethanol treatments; (<b>B</b>) diameter of the zones of inhibition at 8 °C and 15 °C. The error bars are standard deviations of the diameters.</p>
Full article ">Figure 2
<p>Germule suppression assay. (<b>A</b>,<b>B</b>) represent the <span class="html-italic">Pd</span> germination assay for treatments with water, anhydrous ethanol, and various concentrations of propolis at 8 °C and 15 °C, respectively. The white arrowheads indicate the mycelial extension of the <span class="html-italic">Pd</span> spores at two different incubation temperatures. The black arrowheads indicate the inhibition of the <span class="html-italic">Pd</span> spores on exposure to propolis at different concentrations. The green arrowheads indicate the formation of white <span class="html-italic">Pd</span> lawns resulting from the treatment of spores with water or anhydrous ethanol.</p>
Full article ">Figure 3
<p>Micrographs of <span class="html-italic">Pd</span> spores displayed at 10× and 40× magnification: (<b>i</b>–<b>ii</b>) elliptical shape of untreated <span class="html-italic">Pd</span> spores; (<b>iii</b>–<b>iv</b>) deformed <span class="html-italic">Pd</span> spores treated with 1% propolis.</p>
Full article ">Figure 4
<p>(<b>A</b>) MALDI-TOF-MS of a propolis sample at a mass range of 100–400 Da. Each of the peaks on the mass spectrum represents a distinctive compound in our propolis sample. The numbers above the peaks correspond to the compounds listed in <a href="#antibiotics-07-00002-t001" class="html-table">Table 1</a>; (<b>B</b>) magnified version of the mass spectrum at a mass range of the 100–225 Da.</p>
Full article ">
2331 KiB  
Article
Sperm Quality during Storage Is Not Affected by the Presence of Antibiotics in EquiPlus Semen Extender but Is Improved by Single Layer Centrifugation
by Ziyad Al-Kass, Joachim Spergser, Christine Aurich, Juliane Kuhl, Kathrin Schmidt, Anders Johannisson and Jane M. Morrell
Antibiotics 2018, 7(1), 1; https://doi.org/10.3390/antibiotics7010001 - 21 Dec 2017
Cited by 12 | Viewed by 7074
Abstract
Contamination of semen with bacteria arises during semen collection and handling. This bacterial contamination is typically controlled by adding antibiotics to semen extenders but intensive usage of antibiotics can lead to the development of bacterial resistance and may be detrimental to sperm quality. [...] Read more.
Contamination of semen with bacteria arises during semen collection and handling. This bacterial contamination is typically controlled by adding antibiotics to semen extenders but intensive usage of antibiotics can lead to the development of bacterial resistance and may be detrimental to sperm quality. The objective of this study was to determine the effects of antibiotics in a semen extender on sperm quality and to investigate the effects of removal of bacteria by modified Single Layer Centrifugation (MSLC) through a colloid. Semen was collected from six adult pony stallions (three ejaculates per male). Aliquots of extended semen were used for MSLC with Equicoll, resulting in four treatment groups: control and MSLC in extender with antibiotics (CA and SA, respectively); control and MSLC in extender without antibiotics (CW and SW, respectively). Sperm motility, membrane integrity, mitochondrial membrane potential and chromatin integrity were evaluated daily by computer-assisted sperm analysis (CASA) and flow cytometry. There were no differences in sperm quality between CA and CW, or between SA and SW, although progressive motility was negatively correlated to total bacterial counts at 0 h. However, MSLC groups showed higher mean total motility (P < 0.001), progressive motility (P < 0.05), membrane integrity (P < 0.0001) and mitochondrial membrane potential (P < 0.05), as well as better chromatin integrity (P < 0.05), than controls. Sperm quality remained higher in the MSLC groups than controls throughout storage. These results indicate that sperm quality was not adversely affected by the presence of antibiotics but was improved considerably by MSLC. Full article
(This article belongs to the Special Issue Top 35 of Antibiotics Travel Awards 2017)
Show Figures

Figure 1

Figure 1
<p>Modified Single Layer Centrifugation (MSLC). A 50 mL sterile tube with a sterile 5 mL plastic inner tube inserted through the middle of the cover, and a small hole that had previously been made at the edge of the lid. The colloid is poured into the 50 mL tube, the lid containing the insert is screwed on, and the semen samples are added through the small hole near the edge of the lid. After centrifugation, the sperm pellet can be retrieved easily by passing a long Pasteur pipette through the plastic insert.</p>
Full article ">Figure 2
<p>Total motility in control and MSLC samples, with and without antibiotics, during storage for 96 h at 6 °C. Values are Least Squares Means ± SE (<span class="html-italic">n</span> = 18). Note: * P &lt; 0.05, ** P &lt; 0.01, **** P &lt; 0.0001. Abbreviations: CA, control with antibiotics; SA, modified single layer centrifugation with antibiotics; CW, control without antibiotics; SW, modified single layer centrifugation without antibiotics.</p>
Full article ">Figure 3
<p>Membrane integrity in control and MSLC samples, with and without antibiotics, during storage for 96 h at 6 °C. Values are Least Squares Means ± SE (<span class="html-italic">n</span> = 18). Note: * P &lt; 0.05, *** P &lt; 0.001, **** P &lt; 0.0001; not possible to analyze samples at 0 h. Abbreviations: CA, control with antibiotics; SA, modified single layer centrifugation with antibiotics; CW, control without antibiotics; SW, modified single layer centrifugation without antibiotics.</p>
Full article ">Figure 4
<p>DNA fragmentation index for control and MSLC samples, with and without antibiotics (Least squares means ± SE; <span class="html-italic">n</span> = 18). Note: * P &lt; 0.05; it was not possible to analyze samples at 0 h. Abbreviations: CA, control with antibiotics; SA, modified single layer centrifugation with antibiotics; CW, control without antibiotics; SW, modified single layer centrifugation without antibiotics.</p>
Full article ">Figure 5
<p>Total bacterial colony counts per treatment group relative to control with antibiotics (at 0 h); control with antibiotics has been normalized to 1 arbitrary unit. Abbreviations: CA, control with antibiotics; SA, modified single layer centrifugation with antibiotics; CW, control without antibiotics; SW, modified single layer centrifugation without antibiotics.</p>
Full article ">Figure 6
<p>Correlation between total bacterial colony counts per sample and progressive motility at 0 h (r = −0.24; P &lt; 0.05).</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop