dedup-isc-ft-v107-score
float64 0.3
1
| uid
stringlengths 32
32
| text
stringlengths 1
17.9k
| paper_id
stringlengths 8
11
| original_image_filename
stringlengths 7
69
|
---|---|---|---|---|
0.414703 | 935384a3b7644a2aa3f173390f08a037 | Satellite classification of Hantavirus Pulmonary Syndrome for the original outbreak region in the U.S. Southwest. Methods are described in [80]. Original landscape analysis is shown using Landsat TM imagery from the 1992 for 1993 outbreak (right). Suitability of the same region (left) four years prior to the recognized outbreak using the same color scale. | PMC10383283 | viruses-15-01461-g003.jpg |
0.42733 | 8c93bf1f27d64e85812a25b6506c3780 | Location of the study fields and survey points. | PMC10383411 | sensors-23-06466-g001.jpg |
0.483762 | c30dfb7d2729480db7fb34be868b56ad | Flowchart of (a) rice lodging assessment and (b) soil fertility estimation. UAS: unmanned aircraft system; DSM: digital surface model; CHM: canopy height model. | PMC10383411 | sensors-23-06466-g002.jpg |
0.439778 | 29962795f090449191b9e40db4f398d8 | Comparison between (a) rice growth status assessed by the NDVI, (b) estimated SAN distribution, and (c,d) lodging severity. NDVI: normalized difference vegetation index; SAN: available nitrogen in soil; CHM: canopy height model. | PMC10383411 | sensors-23-06466-g003.jpg |
0.459677 | 8afdf5c5ee424aa284b431c3726d9312 | Variation in rice plant height during the research period (June–September 2019). r: maximum plant height; y: harvest plant height. | PMC10383411 | sensors-23-06466-g004.jpg |
0.440017 | 5e19f803d96548e59adfea345b187722 | (a) Bare soil DTM, (b,c) calculated maximum plant height and harvest plant height using DSM. DTM: digital terrain model. | PMC10383411 | sensors-23-06466-g005.jpg |
0.440503 | 0c23a7f013704c66b828f650d350e4c3 | Correlation between measured plant height and the CHM. | PMC10383411 | sensors-23-06466-g006.jpg |
0.424895 | 7c7a2b7c3d27430f9607b402a91f1a02 | Relationship between SAN and DN of selected explanatory variables with the SAN estimating equation. DN: digital number; CC: correlation coefficient; R2: root mean square; AIC: Akaike’s Information Criterion. | PMC10383411 | sensors-23-06466-g007.jpg |
0.372356 | bdfb1cc62367467283d79c2a47b61239 | Correlation between SAN and inclination angle of the rice plants at harvest (n = 3781). The correlation equation was estimated from inside mesh data (n = 2906). | PMC10383411 | sensors-23-06466-g008.jpg |
0.386551 | cf680b2dcf6849ee9c9d6ea853ea7853 | Subplot of red and NIR bands with soil line. | PMC10383411 | sensors-23-06466-g009.jpg |
0.43225 | 3c9b5e1df5584ed4b8c83a5860198d55 | Visual representation of Crohn’s disease (left) and ulcerative colitis (right). It can be seen some of the usual areas involved in UC and CD. It should be noted that there is substantial variation among patients. | PMC10383667 | medicina-59-01218-g001.jpg |
0.434801 | 7b6d767d35464a7a8aa2207f7eabc6e8 | Schematic representation of the interaction between genetic predisposition and environmental factors in ulcerative colitis (UC) and Crohn’s disease (CD). IBD, in both of its main forms, is likely caused by a combination of underlying genetic conditions and environmental conditions. | PMC10383667 | medicina-59-01218-g002.jpg |
0.469874 | f68ba655faaf4357801ab41de55d1d96 | Histogram describing the age of the patients. The range is from 19 to 82 years old. | PMC10383667 | medicina-59-01218-g003.jpg |
0.417145 | 2ba7ca08bed04f8e8254dd5bf1dbabd8 | Accuracy of the neural network model for a range of number of artificial neurons. No model has an accuracy below 70% or higher than 80.35%. | PMC10383667 | medicina-59-01218-g004.jpg |
0.491794 | 2fdcd2eda81640bab7695505c8d588e8 | In vitro infection of (A) L. (L.) infantum LD strain and (B) the ME clinical isolate. BMDMs were infected with stationary-phase promastigotes at 37 °C and 5% CO2 on coverslips in 24-well plates. After 3 h, the wells were washed to remove non-internalized parasites, and the plate was incubated further under the same conditions. After 72 h, infected macrophages were fixed in methanol (Sigma-Aldrich), stained by the Panoptic hematological method (Laborclin, Brazil), then visualized on a light microscope. Bar: 10 μm. | PMC10383904 | tropicalmed-08-00354-g001.jpg |
0.432835 | f705d84b5ff34d239f72a3690784e032 | Investigation of the MSL in the L. (L.) infantum LD strain and ME clinical isolate, using PCR protocols previously described by Carnielli et al. [11]. (A) Schematic representation of the MSL of approximately 14 kb. In the absence of the MSL, a 1.2 kb DNA fragment is amplified. (B) PCR for the presence (~14 kb) or absence (~1.2 kb) of the MSL in the LD strain (lane 1) and the ME isolate (lane 2). (C) PCR amplification of the MSL genes LinJ.31.2370 (that encodes NUC1) and LinJ.31.2400 (that encodes 3,2-trans-enoyl-CoA isomerase). The size of amplified PCR products is indicated above the figure. As a control, all genomic DNAs were also evaluated by a PCR protocol that amplifies a 1.28 kb product of the hsp70 gene, as previously described by Montalvo et al. [23]. MW—molecular weight in kilobase (kb); (−) negative control for PCR (absence of genomic DNA); (+) positive control for PCR (genomic DNA of L. (L.) donovani DD8 strain); (1) L. (L.) infantum LD strain; (2) L. (L.) infantum ME clinical isolate. | PMC10383904 | tropicalmed-08-00354-g002.jpg |
0.45519 | 6d92de4ac5464bca8e12b2223a5ff339 | The organization of Apis mellifera solinvivirus-1 genomic RNA. (a) The schematic representation of AmSV1 genomic RNA (GenBank accession number OQ540582). The position of the main ORF and putative protein domains is shown. Amino acid positions of the protein domains are indicated above the ORF. Non-structural proteins: Hel—helicase, Prot—3C protease, structural proteins: JR—jelly roll domain of structural viral protein (VP)1, VP2; (b) NGS coverage of the AmSV1 genome. Genetic diversity of AmSV1 apiary-level population used for NGS analysis: (c) Shannon’s diversity profile (sliding window average for 100 nt positions); (d) distribution of polymorphic nucleotides (n = 419) and amino acids (n = 63), showing an alternate allele exceeding 3% in frequency in the apiary-level NGS library. (e) The maximum likelihood phylogenetic tree was generated based on the full-length protein sequences of AmSV1, as well as classified (marked with asterisk) and putative solinviviruses. For SINV3 and RAAV, the sequences of -1 translational frameshift proteins (ORF1-ORF2 fusions) were used. Bootstrap values above 50%, generated from 1000 replications, are shown to the left of corresponding nodes. The bar indicates a 10% sequence difference. | PMC10384192 | viruses-15-01597-g001.jpg |
0.455407 | 32aba2f5238d4ffab26e01e95c9082a9 | Accumulation of AmSV1 in different body parts of individual adult honeybee workers from AmSV1 positive apiaries. (a) Sections of a frozen worker honey bee used for RNA extraction. (b) AmSV1 loads in the head, thorax, and abdomen of 16 worker bees. (c) Correlation between AmSV1 loads in head, thorax, and abdomen. | PMC10384192 | viruses-15-01597-g002.jpg |
0.458101 | 8944849037ec486f9f8492abf0ea1345 | Pupal injection experiment. (a) The quantification of AmSV1 genome equivalents (GE) in the pupae, injected with partially purified AmSV1 preparation (AmSV1) or buffer control (PBS), which were sampled immediately after injection (Time 0) and after 3 days of incubation at +33 °C (3 dpi). Quantification was carried out by qPCR using cDNA-generated random primers, allowing the detection of AmSV1 RNA of both polarities. Dots indicate levels of AmSV1 in individual pupae. Significantly different levels of AmSV1 RNA are indicated by different red letters (ANOVA p < 0.01). (b) Specific detection of negative-strand RNA, a virus replicative intermediate, in virus preparation (lane 0) used for injection. The pupae was injected with partially purified AmSV1 preparation (AmSV1) and sampled immediately after injection, Time 0 (lanes 2 and 3, individual pupae), or after 3 day incubation at +33 °C, with 3 dpi (lanes 6 and 7, two pools of 2 pupae). Control, buffer-injected pupae—PBS—were sampled 3 days after injection (lanes 4 and 5, two pools of two pupae). M, DNA ladder, base pairs (bp). The cDNA was produced using tagged forward primer, PCR amplification was carried out with the primer that was identical to the tag and reverse primer. The arrow marks the position of the expected 141 bp RT-PCR product. | PMC10384192 | viruses-15-01597-g003.jpg |
0.404153 | 7ec481dec5bb4a78ab21c7fe2e17de05 | Spatio-temporal distribution of AmSV1 in the USA. (a) The distribution of AmSV1 in the US apiaries in 2010, 2014, and 2021. (b) Monthly loads (0 = not detected, below 2.8 log10 GE/bee) and (c) monthly distribution of AmSV1 prevalence and loads for the year 2021. | PMC10384192 | viruses-15-01597-g004.jpg |
0.45819 | 5104ccaf24dc4c3cbf114af174df406b | Connections between the prevalence of AmSV1 and (a) other honey bee parasites, including (b) field apiary observations in the U.S. 2021 apiaries. In total, 794 apiary-level samples were tested. Dashed line at OR = 1 represents the null hypothesis that AmSV1 does not associate with listed measures. DWV–A—deformed wing virus, type A; DWV-B deformed wing virus type B; ABPV—acute bee paralysis virus; CBPV—chronic bee paralysis virus; IAPV—Israeli acute paralysis virus; KBV—Kashmir bee virus; LSV2—Lake Sinai virus; Varroa—ectoparasitic mite Varroa destructor; Nosema—Vairimorpha ceranae. EFB—European foulbrood (caused by bacterium Melissococcus plutonius), Sacbrood–caused by sacbrood virus, Chalkbrood–fungal disease of honey bee brood caused by fungus Ascosphaera apis, PMS—Parasitic Mite Syndrome (caused by the mite Varroa destructor), Deformed Wings–could be caused by DWV, Shiny Black—hairless bees, SHB—infestation with small hive beetle (Aethina tumida), Wax Moth—infestation with wax moth (Galleria mellonella), Queen Cells presence, Drone Layer–queen lays unfertilized drone eggs, Queenless—queen is absent in at least one of sampled colonies, Any Queen Issues—combined Queen Cells, Drone Layer, and Queenless. A significant p-value is marked with an asterisk. | PMC10384192 | viruses-15-01597-g005.jpg |
0.518339 | 60a949d762134840901363acc3edb77a | Synthesis and crystal structure of 1·xSol with 50% thermal ellipsoids. All hydrogen atoms except for those of the N−H···O bonds are omitted for clarity. | PMC10384712 | molecules-28-05513-g001.jpg |
0.446461 | f0e7dbed5dbd4c1f93192b28429c6199 | (a) The hydrogen bonds (dashed lines) in 1·xSol. (b) Packing of molecular arrays in 1·xSol viewing along the c axis. All hydrogen atoms except for those of the N−H···O and O−H···N bonds are omitted. | PMC10384712 | molecules-28-05513-g002.jpg |
0.49803 | e7a55666661543b18eef8ba7f56a54fb | (a) Interconversions and (b) PXRD patterns of 1·xSol, 1·2MeOH, 1, and 1R. | PMC10384712 | molecules-28-05513-g003.jpg |
0.42767 | 6e5afbe7311f4a48a817bf2223432377 | (Left) Excitation (dotted lines) and emission (solid lines) spectra of 1·xSol, 1·2MeOH, 1, and 1R and (right) photos of these compounds under 365 nm excitation. | PMC10384712 | molecules-28-05513-g004.jpg |
0.512074 | fd81a398261f40d78c9543db0fc33da9 | Temperature-dependent emission spectra of 1·2MeOH from 80 K to 360 K with 40 K temperature interval (λex = 373 nm). | PMC10384712 | molecules-28-05513-g005.jpg |
0.433536 | b3d6cdae07e84277ae3d128845a3d1e2 | The distribution of HOMOs and LUMOs in 1·xSol. | PMC10384712 | molecules-28-05513-g006.jpg |
0.427927 | 0252a1185b4c4a93bc093161f9f07028 | (Left) Emission spectra of 1·2MeOH, 1G, and 1GR under 367 nm excitation. (Right) Photos of 1G and 1GR under 365 nm LED irradiation. | PMC10384712 | molecules-28-05513-g007.jpg |
0.466745 | 687e8bac6fe749958d595bcb8d903317 | The emission spectra of 1G upon exposure to solvent vapors (PE: petroleum ether; MeCN: acetonitrile; EA: ethyl acetate; PrOH: propanol; iPrOH: iso-propanol). | PMC10384712 | molecules-28-05513-g008.jpg |
0.422579 | f8651014dab84f558a7c255d84c5d8b6 | (left) Emission spectra of 1G after exposure to mixed MeOH/H2O vapors of different MeOH content (λex = 367 nm). (right) Test papers under natural daylight and under 365 nm UV light after exposure to MeOH/H2O vapors. | PMC10384712 | molecules-28-05513-g009.jpg |
0.50165 | 9ae84e49e0714e33a4e7ad935315db44 | (a) XRD patterns; (b) FTIR spectra; and (c) Typical SEM image and EDS analysis of as-prepared samples. | PMC10386041 | materials-16-05027-g001.jpg |
0.474072 | 2b27aa8e16564dcbab5b12227dda4992 | (a) The effects of different loading amounts of HC in ZnBi-LDO and (b) Rate constant k values of all the photocatalysts after 135 min of visible light exposure (catalyst: 1.0 g/L, 2,4-D: 30 mg/L at pH 4.0). | PMC10386041 | materials-16-05027-g002.jpg |
0.491489 | 1735089ceb0c46b185406db090968ddb | (a) PL emission spectra for the ZnBi-LDO and 2% HC-ZnBi-LDO; (b) UV−Vis diffuse reflectance spectra and band gap calculation using Tauc’s plot of as-prepared samples (inset) and (c) XPS spectra of the ZnBi-LDO and 2% HC-ZnBi-LDO (Zn2p, Bi4f, O1s, and C1s spectra). | PMC10386041 | materials-16-05027-g003a.jpg |
0.419076 | 095f5c6b116e4acca2fd2a02c5b736bc | (a) The effects of pH solution on degradation of 2,4-D over 2% HC-ZnBi-LDO photocatalyst; (b) Point of zero charge (pHPZC) of 2% HC-ZnBi-LDO photocatalyst (ΔpH versus pH initial); (c) The effects of the amount of 2% HC-ZnBi-LDO photocatalyst on degradation of 2,4-D and (d) The effects of initial 2,4-D concentration on degradation of 2,4-D over 2% HC-ZnBi-LDO photocatalyst after 135 min of visible light exposure. | PMC10386041 | materials-16-05027-g004a.jpg |
0.480225 | 716108d66bd7480db73cccdfeedf11fe | (a) The effects of various scavengers (tert-butanol, Na2EDTA and p-benzoquinone) on the photodegradation of 2,4-D over 2% HC-ZnBi-LDO after 135 min of visible light exposure and (b) Schematic diagram of the photoinduced e−—h+ separation and transfer of photoinduced e−at the visible light-driven 2% HC-ZnBi-LDO interface. | PMC10386041 | materials-16-05027-g005.jpg |
0.427593 | 3ce6e190c6024e048d3acc71d188f60c | Overview of a test day. | PMC10386048 | nutrients-15-03196-g001.jpg |
0.476508 | a16f2af96bcf4016b380d83220f91ffc | Flowchart diagram of study participant selection and inclusion. | PMC10386048 | nutrients-15-03196-g002.jpg |
0.486019 | 6a9d5ff44e48404586bca2d5ee3cab3d | Postprandial AA levels per protein shake fitted per individual. For three individuals (3, 5, and 10), not all individual curves could be fitted based on absolute values (raw data, see Figure A1). | PMC10386048 | nutrients-15-03196-g003.jpg |
0.463631 | 2cceb46233dc40e4b9e4db42f1476624 | Confidence intervals for comparisons in iAUC, peak height, and time-2-max of TAA and TEAA between the PP and WP interventions on the one hand and the reference BRP on the other. For iAUC, the comparison is the ratio between the iAUC values (dimensionless); for the peak height and time-2-max, the comparisons are given by the absolute differences (µM or min, respectively). Red bars indicate a statistically significant difference (p < 0.05). | PMC10386048 | nutrients-15-03196-g004.jpg |
0.359727 | 608992b829a94ce0acd384edb96dbaf5 | Confidence intervals for comparisons in iAUC and a peak height of individual EAA between the PP and WP interventions on the one hand and BRP as a reference on the other. For iAUC, the comparison is the ratio between the iAUC values (dimensionless); for the peak height, the comparison is given by the absolute differences (µM). Red bars indicate a statistically significant difference (p < 0.05), black bars indicate no statistically significant difference (p > 0.05). | PMC10386048 | nutrients-15-03196-g005.jpg |
0.467389 | 791daad1e83148888a43ae2ab5a02d0d | Comparison of the relative abundance of EAA in the protein (expressed in %) product against the iAUC response for each individual in blood after consumption of the protein source. | PMC10386048 | nutrients-15-03196-g006.jpg |
0.449864 | 5aee345546774343bcd2347487bec0b6 | Postprandial glucose and insulin responses after BRP, PP, or WP shake consumption. Data show averages and SD across participants at each time point. | PMC10386048 | nutrients-15-03196-g007.jpg |
0.464009 | ea8038e8eed543f99d6f96731af9d5c1 | Raw postprandial AA levels per Barley/Rice protein (BRP), pea protein (PP), and whey protein (WP) protein per individual. | PMC10386048 | nutrients-15-03196-g0A1.jpg |
0.446684 | 9c84e034e92d481998fad23c2c103b3c | Confidence intervals for all comparisons of individual amino acids, after imputation compared to BRP. In cases where no peak was observed, data for the incremental area under the curve (iAUC) peak heights were imputed from the raw data without curve fitting. For the iAUC, the comparison is the ratio between the iAUC values of PP and WP intervention and the BRP intervention; for the peak height and time-2-max the comparison is given by the absolute differences (µM or min). Red bars indicate significant differences (p < 0.05) compared to the BRP intervention, black bars indicate no statistically significant difference (p > 0.05). | PMC10386048 | nutrients-15-03196-g0A2.jpg |
0.400049 | 8a95b5e0875845e7a88f4e0b00d896e8 | Long-term variations in filamentous Mn particle densities and DO concentrations at a depth of 90 m at the study site. Vertical lines represent complete overturns of the water column. | PMC10386369 | microorganisms-11-01814-g001.jpg |
0.422406 | d599098d500d49c595b672cb070ab528 | TEM image of filamentous Mn oxide particle collected from BIWAKO-01 culture at day 21. The high-magnification image shows that the filaments have a sheet-type structure. Bar: 100 nm. | PMC10386369 | microorganisms-11-01814-g002.jpg |
0.422803 | ab3ca2f6a7b24fc89a7259ba8f4aae76 | Microscopic images of filamentous Mn oxide particles that formed in laboratory cultures of Bosea sp. BIWAKO-01. The bacterial strain was grown in the presence of green algae: S. dorsidentiferum intact (a) and mucilage-sheath-removed (b) cells; S. arctiscon intact (c) and mucilage-sheath-removed (d) cells. Images were obtained from India-ink-stained specimens. The arrowheads indicate filamentous Mn particles. Bar: 30 μm. | PMC10386369 | microorganisms-11-01814-g003.jpg |
0.396608 | e4e4fa14a35744a78893b226b0a2f1b6 | Microscopic images of filamentous Mn oxide particles collected at a depth of 90 m at the study site. (a) India-ink-stained aggregates including filamentous Mn particles (arrowhead 1), dead algal cells (arrowhead 2), and gelatinous substances (arrowhead 3). Bar: 20 μm. (b) Unstained aggregates including numerous Mn particles and dead algal cells (arrowheads 4 and 5). Bar: 100 μm. (c,d) Differential interference contrast and epifluorescence images of aggregates stained with fluorescein-conjugated LEL. Gelatinous substances are indicated with arrowheads. Bar: 10 μm. | PMC10386369 | microorganisms-11-01814-g004.jpg |
0.458842 | 6092e8f0464249dc9447a605bde0fa6c | Seasonal and vertical variations in Chl.a (a) and total polysaccharide (b) concentrations and filamentous Mn particle densities (c) at the study site. | PMC10386369 | microorganisms-11-01814-g005.jpg |
0.429376 | e42890d61607438b9d14c728c7e88296 | Correlations between the total polysaccharide and Chl.a concentrations (a), total phytoplankton biovolume (b), Chlorophyceae biovolume (c), Cyanobacteria biovolume (d), Bacillariophyceae biovolume (e), and Chrysophyceae biovolume (f). All the data were obtained at a depth of 0.5 m at the study site of Lake Biwa. | PMC10386369 | microorganisms-11-01814-g006.jpg |
0.456221 | f8959c0f0ed24fbcb6e59abeeaf12f1b | PCA biplot for the annual average values of filamentous Mn particle densities and water quality data (a) and biovolume of algal species (b) collected over 18 years at the study site. | PMC10386369 | microorganisms-11-01814-g007.jpg |
0.423642 | 166d2f116cfa46d783509875c437edea | Correlations between the filamentous Mn particle density at 90 m and total phytoplankton biovolume (a) and Chlorophyceae biovolume (b) at 0.5 m. The plots represent the annual average data collected for 18 years. | PMC10386369 | microorganisms-11-01814-g008.jpg |
0.40478 | 062029dff1a9428386d14e55fe6b6e94 | Gametocytes of different Haemoproteus majoris lineages found in blood films: hCCF5 (A–D) in Fringilla coelebs, hCWT4 (E–H) in Sylvia curruca, hPARUS1 (I–L) in Cyanistes caeruleus, hPHSIB1 (M–P) in Phoenicurus ochruros, and hWW2 (Q–T) in Phylloscopus trochilus. Note that gametocytes of all lineages are morphologically indistinguishable; the fully grown gametocytes (C,D,G,H,K,L,O,P,S,T) of all lineages reach the poles of erythrocytes but do not completely encircle the erythrocyte nuclei, which were displaced laterally. Pigment granules were similar in size, form, and number in gametocytes of all lineages. Long arrow—gametocyte nucleus; short arrow—erythrocyte nucleus; arrowhead—pigment granules. Scale bar: 10 μm. | PMC10386383 | pathogens-12-00898-g001.jpg |
0.467853 | 976b4689ea2743f889aa48577ee4d132 | Megalomeronts of different Haemoproteus majoris lineages: hCCF5 (A–D), hCWT4 (E–I), hPARUS1 (J–N), hPHSIB1 (O–S), and hWW2 (T–X) in haematoxylin and eosin (H&E)-stained sections and their corresponding images after chromogenic in situ hybridization (CISH) treatment (inserts and (D,I). Megalomeronts were found in pancreas (A,D), gizzard (B,E–I,M,W,X), lungs (C,K,V), kidneys (J,N–S), liver (L), muscles (T), and heart (U) of their host. Note the variously shaped, interconnected cytomeres in developing megalomeronts (A–C,J,M) and the more densely aggregated and connected cytomeres in megalomeronts of hCWT4 and hWW2 (F,G,T,W,X). Very young megalomeronts (D,I,N,S) were found in the infections of four lineages. The host cell nucleus was slightly enlarged and visible in the very young megalomeronts (D,I,N,S) but absent in more developed (A–C,E–H,J,K,M,O–P,T–X) and ruptured (L) megalomeronts. Ruptured megalomeronts (L) were found in the liver of one individual. One megalomeront was found to appear in serial sections (O–R) showing how different its morphology and size can be depending on the analyzed section. Megalomeronts were found solitary in the tissues, and sometimes several megalomeronts were found in the same section located close to each other (A,T). Inflammatory reactions were observed around several megalomeronts (B,F,G,J–M,O–R,T,U). Megalomeronts were surrounded by a thick capsular-like wall, except for the very young ones. Cytomeres were readily visible in stages of advanced development. Long arrow: megalomeront; Short arrow: capsular-like wall; Cross: inflammatory reaction; Arrowhead: enlarged host cell nucleus. Scale bar: 100 μm (A,T); 25 μm (B–S,U–X). | PMC10386383 | pathogens-12-00898-g002.jpg |
0.485103 | 5478c73af4cb48e59b8fa97e99a2ff73 | Exo-erythrocytic stages (megalomeronts (A–F), and meronts (G–I)) of different Haemoproteus species in co-infection with H. majoris found in H&E-stained sections and CISH-tested sections (inserts and (C)): co-infections of different H. majoris lineages—hWW2, hPHSIB1, hPARUS1, and hCWT4—present (A) and hWW2 with hPHSIB1 (B–F) present; H. majoris hCCF5 with H. fringillae (unknown lineage) (G); H. majoris hCCF5 with H. fringillae and H. magnus (unknown lineages) (H,I). The tissue stages were found in the brain of a Parus major (A) and a Phoenicurus ochruros (C); the kidneys of a Phylloscopus sibilatrix (B); the intestine (D), gizzard (E), and heart (F) of a Phoenicurus ochruros; and the lungs of Fringilla coelebs (G–I). The developing megalomeronts (A,E,F) were surrounded by a capsular-like wall. The very young megalomeronts (B–D) were seen in cells with still the host cell nucleus present, which was not present in developing megalomeronts (A,E,F). Meronts in the lungs were seen either following the capillaries (G) or grouped tightly together (H,I) in a blood vessel of the lungs of F. coelebs. Cytomeres were readily visible in megalomeronts (F) as well as in developing and maturing meronts (I). The CISH signals were deep purple in the developing exo-erythrocytic stages (inserts and (C)). Long arrow: megalomeront; short arrow: capsular-like wall; arrowhead: host cell nucleus; opened arrowhead: meront. Scale bar: 25 μm (A–D,F–I); 100 μm (E). | PMC10386383 | pathogens-12-00898-g003.jpg |
0.617007 | f48e17473a654296bba06974aeabeb9d | Structure of (a) Reichardt’s Dye # 30 and (b) THPP. | PMC10386554 | molecules-28-05516-g001.jpg |
0.4862 | 89056fd6312146838bd2623eb29cedee | Spectral characteristics of THPP in the Soret band and Q-band region at [OH−] = 0.04 mol/L as a function of the volume percentage of X expressed in (a) XDMF, (b) Xacetone, (c) Xmethanol, and (d) Xacetone: (1) 0, (2) 20, (3) 40, (4) 60, (5) 80, (6) 98. Inset: different variations of λmax of THPP with the volume percentage of X. | PMC10386554 | molecules-28-05516-g002.jpg |
0.411896 | d77314e3066548d895058409df17cc0c | Resonance Raman spectra of THPP in the range of 900–1600 cm−1 in alkaline solutions ([OH−] = 0.04 mol/L) of (A) 100% H2O, (B) 80% DMF-20% H2O, (C) 80% acetone-20% H2O, (D) 80% methanol-20% H2O, (E) 98% acetone-2% H2O, (F) 98% methanol-2% H2O, (G) 80% acetone-20% DMF, and (H) 98% acetone-2% DMF using a 514.5 nm excitation. The small negative peaks are due to solvent subtraction or noise. * = solvent band. | PMC10386554 | molecules-28-05516-g003.jpg |
0.428467 | f4ae04816be44c71a9339d719f508b26 | Relationship between the C-O stretching frequencies of the phenoxide anion substituents and the volume percentage of the organic solvents in three series of the mixed solvents. | PMC10386554 | molecules-28-05516-g004.jpg |
0.470035 | 58f93a083b1b4908834f37b2e4a12d93 | Relationship between the λmax of the n-π* transition band and the corresponding C-O stretching frequency of the phenoxide anion substituents in three series of mixed solvents. | PMC10386554 | molecules-28-05516-g005.jpg |
0.495266 | 807aa8b6306a4c35994a7c6fb3d6c79f | Relationship between λmax and ET(30) value of the solvent in four series of the mixed solvents. The solvent composition at each point of the four curves is expressed as the volume percentage of X (a) XDMF, (b) Xacetone, (c) Xmethanol, and (d) XDMF: (1) 0, (2) 10, (3) 20, (4) 30, (5) 40, (6) 50, (7) 60, (8) 70, (9) 80, (10) 90, (11) 98. | PMC10386554 | molecules-28-05516-g006.jpg |
0.438286 | fbf9fe27f68e438da9a47a4915d086db | Total mucus production from three contact periods in the slug mucosal irritation assay, expressed as a percentage of the initial bodyweight of the slugs. (a) Fillers; (b) mucoadhesive polymers. Unless otherwise stated, sieve fractions from 32–150 µm were used. n = 3, error bars = sd, * = p < 0.05, ** = p < 0.01. | PMC10386645 | pharmaceutics-15-01852-g005.jpg |
0.375279 | 914e8e12a7d0441586c4fe0b856f3eeb | Characterization of microstructure. (A) Neuron density along the M1 cortex. Cortical depth was normalized (0: deepest portion to 1: pial boundary), shown as a blue line; dashed lines indicate approximate layer boundaries. NeuN immunolabeling was used to estimate neuronal density and morphology. Spatial profiles of neuronal density are shown (n = 3 animals per group; bars and lines indicate mean and standard error, respectively; light gray—Control, red—BCNU treated). The right panel shows a control and two experimental animals (scale bar: 200 μm). Colored squares are amplified (250 μm) in (B). (C) Morphological descriptors as a function of depth for the pooled distributions of neurons of all animals, divided by group. Dashed lines delimit Zone I (0.9–0.7 normalized depth), showing enlarged area and roundness; and Zone II (0.5–0.2 normalized depth), showing increased area; box plots of each zone represent the distributions of neuronal area and roundness (p values for between-group Student’s t-tests; Control 605 ± 73, BCNU 687 ± 50 cells). Data from individual animals are illustrated in Supplementary Fig. 1. | PMC10387479 | 41598_2023_38717_Fig1_HTML.jpg |
0.406013 | ca0e71740e6846e5b82133a315c7c456 | GFAP qualitative evaluation in M1. Top panel shows two representative examples, with immunolabels for NeuN (red), and GFAP (green), in one Control (left) and one BCNU-treated (right) animals; scale bar = 150 μm. The right panels show enlarged views of superficial and deep cortical regions labeled with numbers corresponding to the area and group (Control 1,3; BCNU treated 2,4); scale bar = 50 μm. Bottom panels show three different specimens per group, with magnified views of approximately the same locations as in the top-right panels, immunolabeled for GFAP (green); scale bar = 100 μm. | PMC10387479 | 41598_2023_38717_Fig2_HTML.jpg |
0.405826 | 418249e2966b485193ff3d256d444152 | Layer-specific immunofluorescence in M1. (A) In control animals, Necab1 clearly delineates layer IV (dashed lines), with little expression in more superficial layers. Two experimental animals show disorganization of layer IV and more expression in layers II to III. (B) FoxP2 is normally expressed in layer VI, which shows a sharp upper boundary in control animals (dashed line). Experimental animals show blurring of the layer VI boundary and clusters of heterotopic neurons in superficial layers (gray arrowhead). Scale bar and zoom examples: 200 μm. Both schematic representations of the cortex make reference to the location and extent of the images. | PMC10387479 | 41598_2023_38717_Fig3_HTML.jpg |
0.407494 | 20ab0082ea0d4717a28b4d2a59830f7e | Evaluation of calcium time series. (A) Example of the large field-of-view for a [Ca2+]i signal recording (encompassing all cortical layers), overlaid on the Paxinos and Watson atlas (2006) for reference. Red lines represent the pial boundary (top) and gray-white matter interface (bottom). (B) Population activity of one control and one experimental animal, with each row indicating each neuron’s color-coded signal amplitude (ΔF/Fmin) over time. Arrowheads indicate the pilocarpine stimulus. (C) Enlarged view of an example of a [Ca2+]i signal for a control animal indicating the selected time windows before, during, and after pilocarpine stimulation (arrowhead) considered for network analyses (see Fig. 6). | PMC10387479 | 41598_2023_38717_Fig4_HTML.jpg |
0.453723 | 9a819eb340b44de991b63b40a38b0e99 | Signal power during basal and post-stimulus activity. (A) Cell signal power along the cortical depth during basal activity (left column) and post-pilocarpine (right column); the color scale represents the cell proportion of the total number of neurons evaluated. (B) One-dimensional power distributions (i.e., similar to (A) but across all cortical depths) during basal activity (left) and post-pilocarpine (right). A time-dependent analysis is shown in Supplementary Fig. 2. | PMC10387479 | 41598_2023_38717_Fig5_HTML.jpg |
0.461467 | fae5827ebcba4d7fb0760d95a7fdf127 | Neuron communication before, during, and after the hyperexcitable stimulus. (A) Distribution of group-specific neuron–neuron correlations (Spearman’s ⍴) across time windows (as defined in Fig. 3; pilocarpine stimulus within the second window). (B) Density plots of neuron-wise mean connectivity degree values (k) across the cortical depth and over time windows. (C) Schematic representation of the cortex; curved lines indicate statistically significant reductions of layer-to-layer connectivity in experimental animals (p < 0.05), color-coded according to Cohen’s d. (D) Network metrics over time; no statistically significant between-group differences were identified. | PMC10387479 | 41598_2023_38717_Fig6_HTML.jpg |
0.463372 | 88afd2b4f7a441b8b71052b5ccd044f5 | A 29-year-old healthy man. US (A) and SMI (B) images of the rectus femoris muscle are presented. The RF-CSA value is 2.95 cm2, and the VI value is 0.2. | PMC10388039 | turkjmedsci-53-3-692f1.jpg |
0.510333 | 82dad0f1a5b4497590ab0225382a480c | A 62-year-old woman without CKD or any other chronic disease. US (A) and SMI (B) images show the rectus femoris muscle. The RF-CSA value is 3.65 cm2, and the VI value is 0.4. | PMC10388039 | turkjmedsci-53-3-692f2.jpg |
0.501873 | e976feebf8ec46e5ac16625296ef25c9 | A 58-year-old man with CKD is on hemodialysis. US (A) and SMI (B) images of the rectus femoris muscle are presented. The RF-CSA value is 3.99 cm2, and the VI value is 1.2. | PMC10388039 | turkjmedsci-53-3-692f3.jpg |
0.441914 | f50f9e288edb4feb8a88160baa92a141 | VI values of young, healthy volunteers, adult patients without CKD or another chronic disease, and CKD patients on hemodialysis. The values of all three groups are significantly different from each other. | PMC10388039 | turkjmedsci-53-3-692f4.jpg |
0.361038 | 72ffb8f2fbc14c228e6c48efbe2cef6d | ROC curves with AUC for differentiating CKD patients from adult patients and healthy volunteers. CKD vs. all control group B (A), CKD vs. adult patients group C (B), adult patients group vs. young, healthy group (C). | PMC10388039 | turkjmedsci-53-3-692f5.jpg |
0.430104 | 1d0036b17006460fa9f8bb84d4532e39 | Profile of pro-inflammatory cytokines during the course of the various disease conditions. Scatter plot graphs are plotted showing the concentrations (pg/ml) of A–F IL-1β, IL-6, IL-2, IL-12, IL-17A, GM-CSF and median fluorescence intensities (MFI) of G–V Eotaxin, IFN-γ, RANTES, MCP-1, IL-15, IL-5, IL-1RA, IL-2R, IFN-α, IP-10, TNF, MIG, MIP-1α, MIP-1β, IL-7 and IL-8 in plasma samples collected from uninfected controls (n = 20), patients with severe malaria (n = 27) and uncomplicated malaria (n = 10). For each figure, the interquartile ranges (IQR) are shown as vertical bars and the median is shown as horizontal bars. Significant differences are denoted by *p < 0.05, **p < 0.01, ***p < 0.001, ****p < 0.0001, ns not significant | PMC10388454 | 12936_2023_4652_Fig1_HTML.jpg |
0.52539 | e38808ef7f854079adea63ceeae94f30 | Profile of anti-inflammatory cytokines during the course of the various disease conditions. Scatter plot graphs are plotted showing the concentrations (pg/ml) of A–B IL-4, IL-10 and median fluorescence intensities (MFI) of C IL-13 in plasma samples collected from uninfected controls (n = 20), patients with severe malaria (n = 27) and uncomplicated malaria (n = 10). For each figure, the interquartile ranges (IQR) are shown as vertical bars and the median is shown as horizontal bars. Significant differences are denoted by *p < 0.05, **p < 0.01, ***p < 0.001, ****p < 0.0001, ns not significant | PMC10388454 | 12936_2023_4652_Fig2_HTML.jpg |
0.420745 | a3524d1ab81b48238cceaa97a01eeaf3 | Graphs showing the ROC curves of A–B IL-1β and IL-17A | PMC10388454 | 12936_2023_4652_Fig3_HTML.jpg |
0.462166 | 2bbec3f2107e480ca38e8862883e7732 | Geographical locations of sample collection sites in Rajasthan. (A) National (IND) map. (B) State (RJ) map with sampling district highlighted. (C) Individual sampling locations. | PMC10389044 | fvets-10-1157211-g001.jpg |
0.479934 | ad27a977f1b84dfd91700a505f587d40 | (A) Frequency of livestock animals studied under the present study. (B) Percentage of animals observed as positive using serological and molecular methods. (C) Odds ratio for the likelihood of factors associated with brucellosis. | PMC10389044 | fvets-10-1157211-g002.jpg |
0.441854 | c46d315bb5134d778e088c04bc96bc09 | Flowchart of PRISMA (Preferred Reporting Items for Systematic reviews and Meta-Analyses). | PMC10389420 | js9-109-438-g001.jpg |
0.463368 | 73a6c9bdbc644d43b3d935c32263ed09 | Risk of bias in each study. | PMC10389420 | js9-109-438-g002.jpg |
0.390613 | a0d778db7e5e4c439193ef4e702d8d21 | Risk of bias in general. | PMC10389420 | js9-109-438-g003.jpg |
0.514965 | 8c3274fa0a4741839aa65e50eef15e36 | Funnel plot for the primary outcome. RR, risk ratio; SE, standard error. | PMC10389420 | js9-109-438-g004.jpg |
0.436459 | f4a68db00c014b4586bf8eeab75a3caf | Forest plot for the incidence of postoperative urinary retention between the Tamsulosin group and the Control group. M–H, Mantel–Haenszel. | PMC10389420 | js9-109-438-g005.jpg |
0.489816 | 46c05853fb3e42f986cfa45959f58d8f | Sensitivity analysis for the primary outcome using the leave-one-out analysis. M–H, Mantel–Haenszel. | PMC10389420 | js9-109-438-g006.jpg |
0.478566 | 0f4f543c7e0c41979c1c2af6b58aaf8a | Forest plot for the adverse events. M–H, Mantel–Haenszel. | PMC10389420 | js9-109-438-g007.jpg |
0.492967 | 95a2e95c489e4a0ab1d20c0cbd844d12 | Forest plot for the adverse events after omitting one study. M–H, Mantel–Haenszel. | PMC10389420 | js9-109-438-g008.jpg |
0.491012 | b59b48eff567456c8436d5147156d331 | Forest plot for the incidence of urinary tract infection. M–H, Mantel–Haenszel. | PMC10389420 | js9-109-438-g009.jpg |
0.487684 | 80a0546433584aafaa16d9ec2edbd2a2 | Subgroup analysis based on the mean age of patients for the incidence of postoperative urinary retention between the Tamsulosin group and the Control group. M–H, Mantel–Haenszel. | PMC10389420 | js9-109-438-g010.jpg |
0.428871 | 2b43b355511e43dca81c4461e6ebf1cf | PET-CT axial slice showing both superficial and deep components of the local recurrence with associated malignant pleural effusion from direct extension to the pleural cavity. | PMC10389684 | rjad322f1.jpg |
0.425146 | 5a075d4b83ce406bb1f23e4a26a18bf8 | Flowchart of patient inclusion. | PMC10390104 | turkjmedsci-52-6-1872f1.jpg |
0.498655 | bed7da7c7301433288fc06e62c12256c | Label-free imaging flow cytometry setup used in the experiment. The setup is based on a Mach–Zehnder interferometer and a microfluidic channel, shown in the sketch above. The laser wavefront is illustrated in red. SC, supercontinuum laser source; BS1 and BS2, beam splitters; RR and RR1, retroreflectors; S, sample; MO, microscope objective; M, mirror; TL, tube lens. CMOS, digital camera. | PMC10390541 | 41598_2023_38160_Fig1_HTML.jpg |
0.439467 | e2017002e68c479a9422ad8bee84a526 | Cell image examples for the various cell types acquired. Left column: experimentally captured off-axis holograms, with background holograms shown on its left. Second column: OPD profiles retrieved from the off-axis hologram. Three right columns: semi-synthetic off-axis holograms generated from the coinciding OPD profiles with different augmentations and random \documentclass[12pt]{minimal}
\usepackage{amsmath}
\usepackage{wasysym}
\usepackage{amsfonts}
\usepackage{amssymb}
\usepackage{amsbsy}
\usepackage{mathrsfs}
\usepackage{upgreek}
\setlength{\oddsidemargin}{-69pt}
\begin{document}$$\alpha$$\end{document}α and \documentclass[12pt]{minimal}
\usepackage{amsmath}
\usepackage{wasysym}
\usepackage{amsfonts}
\usepackage{amssymb}
\usepackage{amsbsy}
\usepackage{mathrsfs}
\usepackage{upgreek}
\setlength{\oddsidemargin}{-69pt}
\begin{document}$$\varphi$$\end{document}φ angles, resulting in fringes with various spatial frequencies and directions. These holograms were used for training the deep neural network for cell classification. The white scale bar indicates a length of 10 μm. The color bar on the bottom left refers to the OPD profiles (second column from left) and represents OPD values. | PMC10390541 | 41598_2023_38160_Fig2_HTML.jpg |
0.390097 | 0f232e195b6b47489dd52afc3de24963 | Comparison of classification performance of the SW cell pairs (red), HS cell pairs (green), and WM cell pairs (blue) for CNN classification based on the OPD profiles (solid), and SVM classification based on OPD hand-crafted features (dotted). (a) Comparions of AUC, specificity and recall values. (b) ROC analysis. | PMC10390541 | 41598_2023_38160_Fig3_HTML.jpg |
0.434668 | c4c538e844014bf586100b24d1cb74f2 | Three example frames from the SW-cell flow videos (see Video 1). Left column presents the original holograms; right column presents the coinciding unwrapped OPD profiles. (a) Background hologram without cells. (b) Correct detection of SW480 cells; (c) Misdetection of an SW620 cell in the OPD-based CNN; (d) Correct detection of an SW620 cell, with one misdetection of a cell by the hologram-based CNN; (e) Distribution of the off-axis angles and direction angles during the whole video frames, demonstrating the instability of the off-axis angles during the same experiment. | PMC10390541 | 41598_2023_38160_Fig4_HTML.jpg |
0.430056 | 67a95dd0c1f24ab1ad1ec694a263a747 | Schematic of the enzyme-based UDP-GlcNAc quantification assayThe UDP-GlcNAc quantification method utilizes the enzymatic role of O-GlcNAc transferase (OGT) and BSA-conjugated GlcNAc-acceptor peptide. The UDP-GlcNAc-containing samples react with the GlcNAc-acceptor peptide-BSA and OGT. Following immunodetection with RL2 antibody, O-GlcNAc can be detected. Figure created with BioRender (https://biorender.com/). | PMC10391555 | gr1.jpg |
0.421037 | bced3da3c6b2462ab180ed6b21753f0d | Screening of patient selection and enrollment. LAT: Linezolid-associated thrombocytopenia. | PMC10391562 | gr1.jpg |
0.470451 | f45a8d3a194d4123bcc18e110d204b41 | ADPLCP created by nomogram prediction model was to predict the probability of LAT. To estimate the probability of LAT, mark patient values at each axis, draw a straight line perpendicular to the point axis, and sum the points for all variables. Next, mark the sum on the total point axis and draw a straight line perpendicular to the probability axis. For example, “A” patient was aged 85 years, and his Ccr was 45 mL/min, platelet was 180×109/L, leukocyte was 17.0×109/L, total albumin was 65 g/L, and planned duration of linezolid was 14 days; accordingly, the total point of the patient would be 164, which indicates a probability of 0.73 for developing LAT. ADPLCP: Age, duration, platelet, leukocyte, creatinine clearance, protein; Ccr: Creatinine clearance; LAT: Linezolid-associated thrombocytopenia; TP: Total protein. | PMC10391562 | gr2.jpg |
0.409166 | 7bca3deaaf71497dbf844594a9e0a651 | Calibration curves of nomograms in terms of agreement between the predicted risk and actual observed outcomes. | PMC10391562 | gr3.jpg |
0.446807 | b1b04c5425f04b4eb8eff4debb2df770 | Decision curve analysis of the nomogram for LAT.LAT: Linezolid-associated thrombocytopenia. | PMC10391562 | gr4.jpg |
0.461311 | d948e01077ef4b278433c4091e7edd07 | ROC curve for ADPLCP and other factors. AUC for ADPLCP was 0.802 (95% CI: 0.748–0.856).ADPLCP: Age, duration, platelet, leukocyte, creatinine clearance, protein; AUC: Area under the curve; ROC: Receiver operating characteristic. | PMC10391562 | gr5.jpg |
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.