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
|
---|---|---|---|---|
0.400008 | 4389380010b848459ecb122c741a2a2a | Changes of habitat suitability in 2070 caused by fitness-related and bioclimate variables under a climate warming scenario (RCP 4.5) in five grass lizard (Takydromus) species. Blue, grey and red spots indicate increased, unchanged and decreased suitability by a threshold of 0.01 for average change, respectively. Error bars represent s.d. Suitability change is a value that measures the difference in habitat suitability index at present and under climate change. (Online version in colour.) | PMC9363995 | rspb20221074f06.jpg |
0.497317 | 826305230f9a41798a8b9e65944dcd83 | Flowchart showing the process of study selection. | PMC9364046 | gr1.jpg |
0.456308 | 780376a7a7214f768d718d645b4d1ffe | Images from cancer-focused Molecule of the Month articles.(left) Artistic conception of VegF signaling. VegF (magenta, top left) arrives at the potential site of a new blood vessel by traveling through the blood plasma (tan). VegF brings together two copies of VegFR (top center, lavender/yellow) to form an active dimer. Active VegFR then initiates a signal cascade that leads to intracellular phosphorylation of many proteins, including cadherin (green). The phosphorylated cadherins separate, making room for new blood vessels. Full image available at PDB-101 and in the article on Vascular Endothelial Growth Factor (VegF) and Angiogenesis [14]. (right) Trastuzumab (red/pink) antibody bound to HER2 (blue). The cell membrane is shown schematically in gray. The illustration is built from three PDB structures: the extracellular domain with the antigen-binding fragment (Fab) of trastuzumab (1n8z [15]), the kinase domain inside the cell (3pp0 [16]), and the transmembrane domain (2ks1 [17]). Image available from the article on HER2/neu and Trastuzumab [9]. | PMC9364277 | 41388_2022_2424_Fig1_HTML.jpg |
0.520918 | d78916d3442248abb0b1c4a4d8eb0a3d | PRISMA flow chart for the study. | PMC9364665 | gr1.jpg |
0.494097 | d794d538f1244949aaac80a118b4d6e8 | Productivity by country of publication. | PMC9364665 | gr2.jpg |
0.374525 | 535659d8dc4b443cb729000cb9229b82 | Number of articles selected per year. | PMC9364665 | gr3.jpg |
0.374347 | da856d48548d4594bc176ddb305388c8 | Digital platforms used. | PMC9364665 | gr4.jpg |
0.507485 | 678db273d1694319b9163c51f0fef35e | Keyword network map. | PMC9364665 | gr5.jpg |
0.466356 | c324f302992a4123b592aff81b1186ae | The flow diagram of literature retrieval, screening and exclusion. | PMC9364766 | fimmu-13-967506-g001.jpg |
0.499713 | 3933312121e84725a2faa8a33abccbf2 | Forest plot for effects of tobacco smoking on developing CVD in patients with SLE. systemic lupus erythematosus, SLE; cardiovascular disease, CVD. | PMC9364766 | fimmu-13-967506-g002.jpg |
0.446393 | e6e7bf36ad93449ca0e22c3f5145bded | Subgroup analysis of tobacco smoking on developing CVD in patients with SLE. (A) forest plot according to the study area; (B) forest plot according to different SLE criteria; (C) forest plot according to study quality; (D) forest plot according to proportion of female participants. systemic lupus erythematosus, SLE; cardiovascular disease, CVD. | PMC9364766 | fimmu-13-967506-g003.jpg |
0.459029 | d9880042a5dc4a3f9092929940cdcb5b | GDP per capita, Colombian departments, 2000 vs. 2016 | PMC9365230 | 168_2022_1163_Fig1_HTML.jpg |
0.436042 | 054a49f867064dee8b52f6affc12dbde | GDP/N (unweighted), densities, 2000 vs. 2008 vs. 2016 | PMC9365230 | 168_2022_1163_Fig2_HTML.jpg |
0.514611 | de911cf5c66c4bdbacbfacdfc1258617 | GDP/N (unweighted), ergodic distribution, 2-year transitions Bandwidth: rule of thumb (Silverman 1986) | PMC9365230 | 168_2022_1163_Fig3_HTML.jpg |
0.528359 | dd9dd91f546b457ea958846f7dbc7d33 | GDP/N (unweighted), densities, 2016 (departments and regions).Notes: The vertical lines in each sub-figure represent the normalized GDP per capita for the departments in each region (with the normalization corresponding to \documentclass[12pt]{minimal}
\usepackage{amsmath}
\usepackage{wasysym}
\usepackage{amsfonts}
\usepackage{amssymb}
\usepackage{amsbsy}
\usepackage{mathrsfs}
\usepackage{upgreek}
\setlength{\oddsidemargin}{-69pt}
\begin{document}$$x_{it}=\text {ln}y_{it}/\text {ln}{\bar{y}}_t$$\end{document}xit=lnyit/lny¯t, being \documentclass[12pt]{minimal}
\usepackage{amsmath}
\usepackage{wasysym}
\usepackage{amsfonts}
\usepackage{amssymb}
\usepackage{amsbsy}
\usepackage{mathrsfs}
\usepackage{upgreek}
\setlength{\oddsidemargin}{-69pt}
\begin{document}$$y_{it}$$\end{document}yit the per capita GDP of the department). The labels in each sub-figure refer to the poorest and richest departments in each region | PMC9365230 | 168_2022_1163_Fig4_HTML.jpg |
0.52532 | 10100b3cd8d94dc7b45dd492fa7fee18 | GDP/N (conditioning schemes), densities, 2000 versus 2008 versus 2016 | PMC9365230 | 168_2022_1163_Fig5_HTML.jpg |
0.450691 | 6ac490d401c94594af70641c08abfe82 | GDP/N (conditioning schemes), ergodic distributions, 2-year transitions. Bandwidth: rule of thumb (Silverman 1986) | PMC9365230 | 168_2022_1163_Fig6_HTML.jpg |
0.428101 | d7f4cdd21d4a4500a25925eb2e677edf | GDP/N, densities, unweighted versus population-weighted Bandwidth: local (Loader 1996) | PMC9365230 | 168_2022_1163_Fig7_HTML.jpg |
0.495299 | ef59fe8359294fb0a04951c7a726dd4f | GDP/N, densities, unweighted vs. physically contiguous-conditioned Bandwidth: local (Loader 1996) | PMC9365230 | 168_2022_1163_Fig8_HTML.jpg |
0.499691 | 0872c7553f6f42599d06baad739aef5d | GDP/N, alternative spatially conditioned schemes, densities, 2000 versus 2008 versus 2016 | PMC9365230 | 168_2022_1163_Fig9_HTML.jpg |
0.49437 | e6d4de36ffb241149aa57406e720aa18 | Comparison of the ability of ABFE calculations versus docking to distinguish active compounds from inactives (decoys), shown as Receiver Operating Characteristic (ROC) curves with the Area Under Curve (AUC) statistics for the 30 Tier 1 and 30 Tier 2 compounds of all three protein targets, as labeled. Red: docking results. Blue: ABFE results. These ABFE calculations omit the free energy term for ligand protonation state changes that were incorporated into the docking calculation. However, adding this term to the ABFE results leads to negligible changes in the AUC statistics (maximum change 0.02, mean change 0.00). | PMC9365818 | 41598_2022_17480_Fig1_HTML.jpg |
0.475985 | c24c94d422de433785204385c5816a23 | Comparison of distributions of computed ABFE values for inactive (decoy) and active compounds for Tier 1 and Tier 2 of all three protein targets, as labeled. | PMC9365818 | 41598_2022_17480_Fig2_HTML.jpg |
0.544419 | b849f0308a0248bd9e0af821ed44fa17 | Initial ligand conformations from docking and for the two independent runs of BACE1 active compound CHEMBL1090542. The same initial ligand pose (left) from docking relaxes to two different conformations during the MD equilibration step. In the run that yields the more favorable BFE (middle), the ligand stays close to the initial pose, whereas in the less favorable run (right) the ligand drifts away and loses interaction with the catalytic Asp residues, resulting in much less favorable binding free energy. BACE1 is shown in ribbon representation, the two catalytic Asp residues in ball and stick representation, and the ligand in licorice representation. | PMC9365818 | 41598_2022_17480_Fig3_HTML.jpg |
0.443106 | ebb8c8cb0276453d9bcf8b9e348d3f3b | Copy number variation analysis showed homozygous exon 14 deletion in the ATP6V0A4 (NM_020632) gene (P1). | PMC9366285 | tap-57-4-432_f001.jpg |
0.427055 | 5877391ac99d482a946fff2ba5e0dc57 | Integrative Genomics Viewer data of novel heterozygous c.349+3G>T mutation in the SLC4A1(NM_000342.4) gene (P9) | PMC9366285 | tap-57-4-432_f002.jpg |
0.447719 | b1ca8dcbcf914e89ba7dcd1f23d8da9f | Structure of reported Gαq/11 inhibitors. | PMC9366314 | gr1.jpg |
0.423721 | 22e77b7516944196a8d01a69731f7e8b | Anti-UM efficacy of GQ262 in vivo. Mice were administered with 3 and GQ262 via intraperitoneal injection (ip) for 21 days (n = 6). (A) Tumor volume. (B) Tumor weight. (C) Photo of tumor tissue. (D) Body weight. (E) H&E staining. ∗P < 0.05 vs vehicle; ∗∗∗P < 0.001; ∗∗∗∗P < 0.0001. | PMC9366314 | gr10.jpg |
0.457781 | 4aeb1cd6d74443d59d3bed2374a10edd | Effects of 3 and GQ262 on modulating the Gαq/11 downstream signaling effectors in MP41 xenograft (n = 6). Expression of targeted proteins (ERK, p-ERK, YAP, and p-YAP) was analyzed utilizing Western blot. | PMC9366314 | gr11.jpg |
0.466037 | c8e39657732a444384ed34700dd02a75 | Design and structural optimization strategies for GQ220‒GQ267. The phenyl group (green) is modified to discuss the necessity of the rigid structure. Reduction of the carbonyl (red) and cyclization of the head fragment (blue) are to improve the potency and drug-like property. Replacing the cyclohexylmethyl fragment (pink) is to investigate the significance for inhibitory activity. | PMC9366314 | gr2.jpg |
0.42673 | f072a1bcd0e84e87b1c33421dde856b2 | Summary of structure–activity relationships of imidazopyrazine scaffold derivatives as Gαq/11 inhibitors. | PMC9366314 | gr3.jpg |
0.417289 | f03070d95b224a18b93e15aa89976eae | Cytotoxicity of GQ262. (A) Cell viability after pre-treating with 3 and GQ262 for 2 h. (B) Cell viability after pre-treating with 3 and GQ262 for 24 h. Bar is reported as mean ± SEM (n = 3). | PMC9366314 | gr4.jpg |
0.541648 | 0fa405e037cf42e6b0f6045f3afc136e | Effects of 3 and GQ262 on the BRET signal alterations and Ca2+ release stimulated by agonist. (A) and (B) BRET signals were assessed after adding AngII. (C) The dose–response curves of 3 and GQ262 on AT1R-activated BRET changes after adding AngII. (D) The dose–response curve of AngII-induced Ca2+ release was determined in the presence of 3 or GQ262 at 100 μmol/L. Data is reported as mean ± SEM (n = 3). ∗P < 0.05 vs control; ∗∗∗∗P < 0.0001. | PMC9366314 | gr5.jpg |
0.447222 | e0aea359e21a46848fc13f6cfbd98ce1 | Rescue assay and cellular thermal shift assay (CETSA). (A) MP41 cell viability of GQ262 at different concentrations for 72 h. (B) 92.1 cell viability of GQ262 at different concentrations for 72 h. (C) GQ262 enhanced the thermal stability of Gαq/11 from 37 to 50 °C in MP41 cells. (D) GQ262 enhanced the thermal stability of Gαq/11 from 37 to 50 °C in 92.1 cells. Data is reported as mean ± SEM (n = 3). ∗P < 0.05 vs control; ∗∗P < 0.01; ∗∗∗∗P < 0.0001. | PMC9366314 | gr6.jpg |
0.456108 | 43f3b4417a064b3cb2c09739586f393a | Cell cycle distributions and apoptotic effects of 3 and GQ262 in UM cells. (A) and (B) MP41 cell cycle distributions were measured via flow cytometer after incubating with 3 or GQ262 for 48 h. (C) and (D) Apoptotic MP41 cells were determined after incubating with 3 or GQ262 for 48 h. (E) Analysis of cleaved caspase-7, Bcl-2, and Mcl-1 levels in UM cells. Data is reported as mean ± SEM (n = 3). ∗P < 0.05 vs control; ∗∗P < 0.01; ∗∗∗∗P < 0.0001. | PMC9366314 | gr7.jpg |
0.448797 | 4a1030d6599b41faa7eef958017d4a2c | Inhibition of colony formation of GQ262. (A) Photos of MP41 cell colony formation (magnification, × 40). (B) Numbers of colony formation were counted via Colony-Counter after incubating with 3 or GQ262 for 9 days. (C) Photos of 92.1 cell colony formation (magnification, × 40). (D) Numbers of colony formation were counted via Colony-Counter after incubating with 3 or GQ262 for 9 days. Data is reported as mean ± SEM (n = 3). ∗∗∗∗P < 0.0001 vs control. | PMC9366314 | gr8.jpg |
0.460634 | 521ad6b6c975471aac24c6fe81ca42c6 | Migrative and invasive inhibition of GQ262 on UM cells. (A) Photos of MP41 cell migration (magnification, × 40). (B) Quantitation of cell migration after incubating with 3 or GQ262 for 48 h. (C) Photos of 92.1 cell migration (magnification, × 40). (D) Quantitation of cell migration after incubating with 3 or GQ262 for 48 h. (E) Photos of MP41 cell invasion (magnification, × 100). (F) Quantitation of invading cells. (G) Photos of 92.1 cell invasion (magnification, × 100). (H) Quantitation of invading cells. Data is reported as mean ± SEM (n = 3). ∗∗P < 0.01 vs control; ∗∗∗P < 0.001; ∗∗∗∗P < 0.0001. | PMC9366314 | gr9.jpg |
0.409018 | 071f74313b9c425ebef7bb02ecfb0b74 | Preparation of GQ220‒GQ235. Reagents and conditions: (i) NaBH4, methanol, ice-bath; (ii) DMP, DCM, ice-bath; (iii) glyoxal, NH3/H2O, methanol, rt; (iv) I2, KOH, N,N-dimethylformamide (DMF), rt; (v) ethyl bromoacetate, K2CO3, DMF, rt; (vi) trifluoroacetic acid (TFA), DCM, rt; (vii) K2CO3, DCM/methanol, rt; (viii) BH3·THF, THF, reflux; (ix) HCl, methanol, reflux; (x) Pd (dppf)Cl2, K3PO4, DMF/H2O, 90 °C, Ar; (xi) HATU, DIPEA, DMF, rt. | PMC9366314 | sc1.jpg |
0.485229 | b4a0be3843404eca88532b09f124353f | Preparation of GQ236‒GQ267. Reagents and conditions: (i) 2-bromoacetophenone, Cs2CO3, DMF, rt; (ii) NH4OAc, toluene, reflux; (iii) K2CO3, DMF, rt; (iv) TFA, DCM, 25 °C; (v) K2CO3, CH2Cl2/methanol, rt; (vi) BH3·THF, THF, reflux; (vii) HATU, DIPEA, DMF, rt; (viii) TFA, (i-Pr)3SiH, rt, Ar. | PMC9366314 | sc2.jpg |
0.385177 | 2e5cc191be0c4324940314b3dfcad178 | Interventricular septum velocity vectors in a healthy subject from the apical 4-chamber view at the end of systole (A) and diastole (B). | PMC9366401 | ajc-26-4-269_f001.jpg |
0.466674 | 89b3caeedbdb4282a74433f540a2604a | The longitudinal strain (left corner) and strain rate (right corner) of IVS in a healthy subject during 3 cardiac cycles. Basal segment (A and B; green and blue curves that show IVS right and left sides, respectively). Middle segment (C and D; pink and yellow curves that show IVS right and left sides, respectively) and apical segment (E and F; azure and violet curves that show right and left sides IVS, respectively). IVS, interventricular septum. | PMC9366401 | ajc-26-4-269_f002.jpg |
0.411217 | c0fdfd2fa4234ddbab586cf0edd9590d | Flow diagram for inclusion and exclusion of eligible patients. mRS, modified Rankin Scale; EEG, Electroencephalography. | PMC9366714 | fneur-13-897734-g0001.jpg |
0.525975 | 519153f813844e0fbff02c3e20d17b70 | ROC curves with 50% confidence interval of models and scores for predicting 3-month mortality. ROC curve, receiver operating characteristic curve; QEEG, quantitative EEG; APACHEII, Acute Physiology and Chronic Health Evaluation II; GCS, Glasgow Coma Scale; AUC, area under the curve. The red dots indicate the threshold at which the sensitivity and specificity are best. | PMC9366714 | fneur-13-897734-g0002.jpg |
0.502768 | 168d2027ed55495db0fc11bdd48eaeb8 | Feature contribution of the best model based on all QEEG parameters, APACHEII, and other features. ADR, alpha/delta ratio; BSI, brain symmetry index; APACHEII, Acute Physiology and Chronic Health Evaluation II; Mean AMP, mean amplitude; REG, regularity. | PMC9366714 | fneur-13-897734-g0003.jpg |
0.37182 | 2ea1c11f3e1e48daa229f35251176125 | Feature contribution of the best model based on four QEEG parameters. | PMC9366714 | fneur-13-897734-g0004.jpg |
0.435607 | 956b4dfc319a4ca49dcbec5a8ce49320 | Auditory function parameters in the Fgf22+/+ and Fgf22–/– groups. (A) Mean amplitudes of the ABR wave I in the Fgf22+/+ and Fgf22–/– groups at 8 and 16 kHz. (B) Mean latency values of the ABR wave I in the Fgf22+/+ and Fgf22–/– groups at 8 and 16 kHz. (C) Mean values of the ABR threshold in the Fgf22+/+ and Fgf22–/– groups at 8 and 16 kHz. Data are expressed as the mean ± SD, statistical significance was assessed with two-way ANOVA followed by Bonferroni post hoc test, *P < 0.05, **P < 0.01, ***P < 0.001. | PMC9366910 | fnmol-15-922665-g001.jpg |
0.432828 | 688b9d26f18547efa01e1607a16048a5 | Whole-mount of the cochlea hair cells in Fgf22+/+ and Fgf22–/– groups. (A) Isolated cochlea hair cells were immunostained with phalloidin (red). Scale bars: 20 μm. (B) The bar chart showed no difference in the hair cells between the Fgf22+/+ and Fgf22–/– groups. | PMC9366910 | fnmol-15-922665-g002.jpg |
0.422721 | b60fa6fced53415baea56854f103e125 | The number of synaptic ribbons in Fgf22+/+ and Fgf22–/– mice. (A) Representative cochlear whole-mount preparation (n = 6, middle turn) images of CtBP2-labeled presynaptic ribbons (red) and GluR2-labeled postsynaptic glutamate receptors (green) from the Fgf22+/+ and Fgf22–/– groups showing the overlapped puncta (yellow). Scale bars: 20 μm. (B) The bar chart showed the number of synaptic ribbons per IHC, which was similar in the Fgf22+/+ and Fgf22–/– groups. | PMC9366910 | fnmol-15-922665-g003.jpg |
0.439531 | 552655aecf814c4c943db6438b918048 | The number of synaptic vesicles in the Fgf22+/+ and Fgf22–/– groups. (A) Representative images of synaptic vesicles in the Fgf22+/+ and Fgf22–/– groups under a transmission electron microscopy (TEM). Scale bars: 200 nm. (B) The number of ribbon SVs was significantly reduced in Fgf22–/– mice than in Fgf22+/+ mice. **P < 0.01 vs. the Fgf22–/– group. | PMC9366910 | fnmol-15-922665-g004.jpg |
0.449361 | 3300f7897a7644a2bd5ce1f2a9a37d97 | Changes in Ca2+ current in IHCs of the Fgf22+/+ and Fgf22–/– groups. (A) Representative curves of the Ca2+ current in IHCs of the Fgf22+/+ (black) and Fgf22–/– (gray) groups. The current response was induced by a voltage ramp from –80 to 60 mV and then the leak was subtracted. (B,C) No significance was found in the Ca2+ current amplitude (ICa) and the slope factor (k) between the Fgf22+/+ and Fgf22–/– groups. (D) IHCs from the Fgf22–/– mice (–31.55 mV) have a more negative half-activation voltage (Vhalf) than Fgf22+/+ mice (–28.76 mV). *P = 0.0301, which indicates significant differences with P < 0.05. Statistical significance was assessed with a two-way ANOVA followed by Bonferroni post hoc test. | PMC9366910 | fnmol-15-922665-g005.jpg |
0.443307 | 4a2c145d6fd04cd8ad05fbd87738bbb3 | Changes in exocytosis in IHCs between the Fgf22+/+ and Fgf22–/– groups. (A) Representative Ca2+ currents (ICa) and the resulting capacitance jumps (ΔCm) recorded from IHCs between the Fgf22+/+ (black) and Fgf22–/– (gray) groups. (B,C) ΔCm and the Ca2+ charge (QCa) evoked by stimulations of different durations, from 10 to 200 ms. ΔCm for stimulation of 10 and 30 ms was significantly reduced in the Fgf22–/– group, *P = 0.0301, and P = 0.0251 < 0.05, respectively. (D) The Ca2+ efficiency of triggering exocytosis, assessed based on the ratio of ΔCm/QCa, was reduced significantly for stimulation of 10 and 100 ms in the Fgf22–/– group. Data are expressed as the mean ± SD. *Indicates significant differences with P = 0.0476, and P = 0.0201 < 0.05, respectively. | PMC9366910 | fnmol-15-922665-g006.jpg |
0.515274 | e126acdc8f684cfaacadb30ad5a386f9 | Quantitative real-time PCR analysis in the Corti’s organ of the Fgf22+/+ and Fgf22–/– groups. Fgf22–/– mice displayed downregulation of SNAP-25 and Gipc3 and upregulation of MEF2D. Data are expressed as the mean ± SD, and statistical significance was assessed with a two-way ANOVA followed by Bonferroni post hoc test, *P < 0.05, ***P < 0.001. | PMC9366910 | fnmol-15-922665-g007.jpg |
0.433561 | d7155a2550dd42789ab87f16ad05b13b | Mutations in the TCR signaling pathway. Mutations of TCR signaling-related genes in PTCL. The intracellular pathways after TCR ligation and costimulatory activation were reconstructed from published studies. From left to right: (1) PI3K pathway after CD28/TCR-dependent FYN phosphorylation and ultimately resulting in activation of mTOR and NF-κB pathways; (2) AP-1/MAPK pathway that comprises MALT1-induced JNK activation, and PLCγ1-GRB2/SOS–induced MAPK activation; (3) NF-κB/NFAT pathway proximally initiated by ITK-dependent PLCγ1 activation; and (4) GTPase-dependent pathway, including RHOA, responsible for cytoskeleton remodeling upon costimulatory/TCR activation. Asterisks indicate mutations affecting the respective molecules. | PMC9367541 | cancers-14-03716-g001.jpg |
0.41159 | 250648bb384b4a7884f4a8ce24647751 | Signaling motifs in the cytoplasmic tail of the human CD28 and its binding partners. The human CD28 possesses a 41 amino acid-long cytoplasmic tail that includes three potential protein-protein interaction motifs (highlighted in red). The phospho-Tyr173 within the YMNM motif serves as a docking site for the SH2-containing proteins, p85, GRB2 and GADS. The PRRP motif can interact with the SH3 domain of ITK and LCK. The PYAP motif can interact with the SH3 domain of GRB2, GADS, and LCK. | PMC9367541 | cancers-14-03716-g002.jpg |
0.40151 | 9849bd1b409546689d05cdf073ab6655 | VAV1-mutant proteins resulting from nonsynonymous mutations, in-frame deletions, and fusion with various partners identified in PTCL-NOS, AITL, ALCL, and ATLL. | PMC9367541 | cancers-14-03716-g003.jpg |
0.453016 | ab435c1619684a8d9a93c6c21ca657f8 | Schematic diagram showing structure of the CTLA4-CD28 gene fusion. (A) Schematic diagram of the gene fusion. SP: signal peptide; TM: transmembrane region. (B) In normal T cells, activation of CD28 stimulates proliferation, which is inhibited by CTLA4 signaling. In tumor cells expressing the fusion protein, CTLA4 activation would aberrantly stimulate proliferation through the intracellular CD28 domain. | PMC9367541 | cancers-14-03716-g004.jpg |
0.484957 | 78d5e5ed6fe2485d9ba6aa3623f0e35d | Examples of sufficiency and necessity logic are extracted from [15,16,17], respectively. | PMC9367758 | ijerph-19-09402-g001.jpg |
0.462777 | 15bfcbbe15c7435c83c692f720920739 | Force of extrusion method. | PMC9367761 | foods-11-02244-g001.jpg |
0.447148 | d4ed4355fe284b998048477b1ee79775 | Back extrusion measurements for different proteins at different concentrations. Note: CS (Control Sample), S3 (3% Soy), S5 (5% Soy), C3 (3% Cricket), C5 (5% Cricket), A3 (3% Albumin), and A5 (5% Albumin). Different superscript letters in the same row indicate a significant difference (p < 0.05). | PMC9367761 | foods-11-02244-g002.jpg |
0.37433 | 7e279229136e418b8451174b341287f1 | Force of extrusion measurements for different proteins at different concentrations. a) Consistency b) Firmness Note: CS (Control Sample), S3 (3% Soy), S5 (5% Soy), C3 (3% Cricket), C5 (5% Cricket), A3 (3% Albumin), and A5 (5% Albumin). Different superscript letters indicate a significant difference (p < 0.05). | PMC9367761 | foods-11-02244-g003.jpg |
0.426485 | 81d20ab75cf74ec5b8a8564774864d00 | Pictures of 3D printing formulations with different proteins. Note: CS (Control Sample), S3 (3% Soy), S5 (5% Soy), C3 (3% Cricket), C5 (5% Cricket), A3 (3% Albumin), and A5 (5% Albumin). (A) 8 layers hexagon shape top (B) 8 layers hexagon shape side (C) 13 layers flower shape top (D) 13 layers flower shape side (E) 11 layers flower shape top (F) 11 layers flower shape side (G) 9 layers flower shape top (H) 9 layers flower shape side (I) 28 layers mountain shape top (J) 28 layers mountain shape side. | PMC9367761 | foods-11-02244-g004.jpg |
0.496466 | 44997f8246d843b192d5cf51087e23f7 | Distribution of hand grip strength according to age groups and sex. | PMC9367881 | ijerph-19-09745-g001.jpg |
0.474079 | aef580a833214066adbcb80881b806bf | A framework of the themes and subthemes discussed. | PMC9368609 | ijerph-19-09672-g001.jpg |
0.455305 | 080ae3f1bcbf4a18b195e01adc05a084 | The effects of osteostatin on osteoclast differentiation. Osteoclast precursors were stimulated with M-CSF and RANKL for osteoclast differentiation in the presence or absence of osteostatin (final concentrations: 100, 250 and 500 nM) for 7–9 days. Cells were TRAP and hematoxylin stained, and TRAP+ positive cells with 3 or more nuclei were counted under a light microscope. (A) Representative images; Bar = 100 μm. (B) TRAP+ multinucleated cells (MNCs) per well are expressed as the mean ± S.D. of three independent experiments; ** p < 0.01 vs. M-CSF+RANKL. | PMC9369336 | ijms-23-08551-g001.jpg |
0.448163 | 99ead51f78314792b96772395591f899 | The effects of osteostatin on resorption. Differentiated osteoclasts were seeded on a 96-well osteoassay plate and incubated with RANKL in the presence or absence of osteostatin (final concentrations: 100, 250 and 500 nM) for 2 days. (A) Representative images of resorption pits; Bar = 100 μm. (B) Percentage of resorption area. Values are the mean ± S.D. of four independent experiments. | PMC9369336 | ijms-23-08551-g002.jpg |
0.475019 | 12b738f9d41d4be7af8d2d55a4d31155 | The effects of osteostatin on the mRNA expression of osteoclast markers. (A) Expression levels at 2 days of differentiation. (B) Expression levels at 7 days of differentiation. Osteoclast precursors were incubated with M-CSF and RANKL in the presence or absence of osteostatin (final concentrations: 100, 250 and 500 nM). The levels of mRNA expression were determined with qRT-PCR and normalized to glyceraldehyde-3-phosphate dehydrogenase (GAPDH). Values are the mean ± S.D. of four independent experiments; + p < 0.05, ++ p < 0.01 vs. control; * p < 0.05, ** p < 0.01 vs. M-CSF+RANKL. | PMC9369336 | ijms-23-08551-g003.jpg |
0.392061 | ccd344b0b59b4831a7d46fad30bb16d1 | The effects of osteostatin on NFATc1 nuclear translocation. Osteoclast precursors were stimulated with M-CSF and RANKL in the presence or absence of osteostatin (final concentrations: 100, 250 and 500 nM) for 2 days, and then NFATc1 nuclear translocation was examined by immunofluorescence. (A) Representative images; Bar = 50 μm. Nuclei were stained by DAPI. (B) The nuclear/total integrated optical density (IOD) ratio was obtained to compare NFATc1 nuclear translocation. Results are expressed as the mean ± S.D. of three independent experiments; ++ p < 0.01 vs. control; * p < 0.05 vs. M-CSF+RANKL. | PMC9369336 | ijms-23-08551-g004.jpg |
0.40749 | 110c42545f4943fcb67c46a1548d0bf7 | Morphological images of Bulung Anggur (A,B) and Bulung Boni (C,D). | PMC9370202 | molecules-27-04879-g001.jpg |
0.397664 | 4f89e2802ed14844a2ea4ec4ce65e293 | Chromatogram of Bulung Boni ethanol extract. The largest AUC showed in 14.250 of retention time, indicating that Cyclohexanamine is the dominant chemical of Bulung Boni. | PMC9370202 | molecules-27-04879-g002.jpg |
0.461891 | b465b9bded0540d89b5fdfe502531f14 | Chromatogram of Bulung Anggur ethanol extract. The largest AUC showed in 14.259 of retention time, indicating that Terephthalic acid is the dominant chemical of Bulung Boni. | PMC9370202 | molecules-27-04879-g003.jpg |
0.453086 | 5cc58df97f7c48aaa001a714a63e3d95 | Electrophoresis results of PCR amplification of Bulung Anggur (1) and Bulung Boni (2) with tufA as a primer. | PMC9370202 | molecules-27-04879-g004.jpg |
0.473312 | 02e141986cc14a2ca055264b4cd60830 | Phylogenetic tree for Bulung Boni and Bulung Anggur based on tufA sequences, constructed with maximum likelihood. | PMC9370202 | molecules-27-04879-g005.jpg |
0.413739 | 29c3450864444879aaa77377827275b0 | Map of sampling sites. (A: sampling site of Bulung Anggur; B: sampling site of Bulung Boni). | PMC9370202 | molecules-27-04879-g006.jpg |
0.450813 | c7912098b8024cd68f88e4d94f5ccc76 | (a) Schematic illustration for the structures of GO, partially reduced GO and SrGO; (b) Schematic illustration for the preparation of SrGO membranes. | PMC9370331 | polymers-14-03068-g001.jpg |
0.46252 | 635c9f87afc647298c6d5f62b7127958 | (a) XPS spectrum of a SrGO film; (b) Zeta potentials of a GO film and a SrGO film; (c) C1s spectrum of a GO film; (d) C1s spectrum of a SrGO film. | PMC9370331 | polymers-14-03068-g002.jpg |
0.446821 | 3cde211e5f60493ba3fbd4f7ae8a5c6b | (a) Cross-sectional SEM images of the SrGO membranes; (b) Magnified area marked in a; (c) Magnified area marked in b; (d) Magnified area marked in c. | PMC9370331 | polymers-14-03068-g003.jpg |
0.435542 | 5f91c1b0af05420aa6ce7014b4af5b46 | (a) Permeances of SrGO membranes and GO membranes with various loading amounts; (b) Rejection ratios of SrGO membranes and GO membranes toward various molecules; (c) Permeance and rejection ability toward CR of SrGO membranes with various loading amounts; (d) Performance comparison of the SrGO membranes and other graphene-based membranes in terms of their permeance and rejection ability toward CR (references are shown in Table S2). | PMC9370331 | polymers-14-03068-g004.jpg |
0.429375 | 205a1ae54aa44427ba2d0f6740ab6440 | (a) Rejection ratios of various targets by SrGO membranes at different voltages during filtration of corresponding solutions; (b) Rejection ratios of various targets by SrGO membranes at 0 V and 2.0 V during filtration of their mixed solution. | PMC9370331 | polymers-14-03068-g005.jpg |
0.463446 | 93986a90fb4544d3ba14519fc4a838ed | (a) SEM image of the surface of the SrGO membranes after filtration of copper nitrate solution under electrochemical assistance at 2.0 V; (b) High-resolution SEM image of the area marked in a. | PMC9370331 | polymers-14-03068-g006.jpg |
0.491313 | d7448f93579a474d8565dfa6933d25a6 | Graphical representation of ATR-FTIR spectra of: (a) different polysaccharides; (b) different types of membranes. | PMC9370371 | molecules-27-05026-g001.jpg |
0.506645 | 651ee8387b684054b22b86c0d1bda22e | Graphical representation of UV-Vis spectra of: (a) CH-Ul membrane with different concentrations of ECR; (b) CH-Ul membranes after immersion on 10 mg L−1 Al(III) solution. | PMC9370371 | molecules-27-05026-g002.jpg |
0.446076 | 9bcf14919d754e8d9318f42863b1d884 | Variation in the absorbance as a function of time in different types of membranes. | PMC9370371 | molecules-27-05026-g003.jpg |
0.471294 | 580d1cdaf8c94a96b816d1ad961efaec | Graphical representation of absorption spectra for the different types of sensing membranes in solutions with different Al(III) concentrations: (a) CH-PSil; (b) CH-CS; (c) CH-Fu; (d) CH-Ul. | PMC9370371 | molecules-27-05026-g004.jpg |
0.390835 | f83238f0ee1a4d74ad9efe4c6bb25512 | Calibration curve of the signal for the different membranes in 0.1 to 10 mg L−1 of Al(III). | PMC9370371 | molecules-27-05026-g005.jpg |
0.49129 | 61fad683e589418e8283ba5c6e03fbea | Effect of incorporation of 2 mL CTAB at 1 mmol L−1 into the Al(III) test solution, demonstrated by the variation in (a) peak wavelength and (b) absorbance intensity as a function of concentration of Al(III) solution. | PMC9370371 | molecules-27-05026-g006.jpg |
0.422894 | caef1563173542ecb6c5ea9e9627fe1c | Graphical representation of the UV-Vis spectra and respective image of the CH-CS membrane after immersion in different concentrations of Al(III): (a) without CTAB; (b) with CTAB. | PMC9370371 | molecules-27-05026-g007.jpg |
0.467004 | e400384e10c24c0aa78364781c2f440a | Calibration curves with the different membranes for 0.1 to 3 mg L−1 of Al(III) with CTAB. | PMC9370371 | molecules-27-05026-g008.jpg |
0.418283 | 8760d4aaeca2405e88a65c7ebd123331 | Graphical representation of the UV-Vis spectra of the membrane CH-CS after immersion in different concentrations of: (a) Al(III); (b) Cu(II). | PMC9370371 | molecules-27-05026-g009.jpg |
0.438582 | 11361b9f04ba449babc9e0ed89cafba4 | Selectivity data for the different interfering species. | PMC9370371 | molecules-27-05026-g010.jpg |
0.425943 | 19ca3859c83c47fb9bfe80268dcc0e76 | Graphical representation of the absorption spectra of CH-CS membranes after immersion in samples containing CTAB and: (a) Al(III); (b) Cu(II). | PMC9370371 | molecules-27-05026-g011.jpg |
0.45401 | 4a5329ed214e4ca48efc9f2f1a717226 | Absorption spectra of CS membranes on immersion in CTAB-supplemented solutions of Al(III) and Cu(II) at 20 mg L−1 concentration—comparison between the signals read at 605 nm and 624 nm. | PMC9370371 | molecules-27-05026-g012.jpg |
0.455537 | 37d32538aceb46b4992d883399b395bf | Representation of membrane bending used for determination of its malleability. | PMC9370371 | molecules-27-05026-g013.jpg |
0.435185 | 584d6acdfe3441faa7af81d4f700a8c9 | Image of the solid sample holder utilized for membrane analysis. | PMC9370371 | molecules-27-05026-g014.jpg |
0.444517 | 83480bc00bf347d6980046c819eb0df6 | A typical AEM process in acid media is illustrated schematically. | PMC9370661 | nanomaterials-12-02618-g001.jpg |
0.409123 | c7eb0be6523e48308dc3c3a7335b6a06 | (a) TEM image, (b) HAADF–STEM image, (c) LSV curves of RuCu NSs/C−350 °C, RuCu NPs/C−350 °C, Ir/C, and Pt/C in 0.5 M H2SO4, (d) AFM images and corresponding thickness, (e) high–magnification TEM image, and (f) Reaction pathway of acidic OER on RuCu NSs. Reprinted with permission from Ref. [53], Copyright 2019, John Wiley and Sons. | PMC9370661 | nanomaterials-12-02618-g002.jpg |
0.396301 | 28206e9ff63e4cdda7d1c191e5ad0527 | (a,b) HAADF–STEM image of the Ru3Ni3 NAs. LSV curves of the Ru3Ni3 NAs, Ru3Ni2 NAs, Ru3Ni1 NAs, Ru NAs, and Ir/C under different acidic conditions, (c) in 0.5 M H2SO4, (d) in 0.05 M H2SO4. Life test of the Ru3Ni3 NAs in (e) 0.5 M H2SO4 and (f) 0.05 M H2SO4 solutions at 5 mA cm−2. Reprinted with permission from Ref. [62], Copyright 2019, Cell Press. | PMC9370661 | nanomaterials-12-02618-g003.jpg |
0.457233 | 51b3d94188724a55b9625c4801071f2f | (a) Schematic depiction of the production of E–Ru/Fe ONAs, (b) XRD pattern, and (c) high–magnification TEM picture of P–Ru/Fe NAs, (d) High–magnification TEM picture, (e) HAADF–STEM–EDS elemental mappings, (f) HAADF–STEM image and (g–i) HRTEM images of E-Ru/Fe ONAs. Reprinted with permission from Ref. [61], Copyright 2021, Elsevier. | PMC9370661 | nanomaterials-12-02618-g004.jpg |
0.482574 | d5628974289c44169c76f15af10623e6 | (a) Scheme demonstrating the fabrication of Au–Ir catalysts, (b) XRD patterns of carbon paper and Au–Ir catalysts, (c) TEM image of a representative Au-Ir heterostructured particle formed after OER test, insets are HAADF–STEM image and EDS mapping of the particle, (d) LSV curves of different catalysts in 0.1 M HClO4 solution, (e) Ir–mass–based OER activities of Ir and Au-Ir catalysts, (f) HRTEM image of the region indicated by a yellow dashed square in the particle shown in (c), (g) Electron diffraction pattern of the particle shown in (c), (h) Chronopotentiometric measurements of Ir and Au-Ir catalyst, (i) Overpotentials of different Ir-based motivations at 10 mA/cm2. Reprinted with permission from Ref. [65], Copyright 2021, American Chemical Society. | PMC9370661 | nanomaterials-12-02618-g005.jpg |
0.427621 | 446048f2333e4e47a5a3138ae0839246 | Characterization of the Ir–IrOx/C-20. (a) surface image, (b) cross–section SE, (c,d) Low-magnification HAADF–STEM image. Inset in (d): The nanoparticle size statistics diagram, (e,f) Spherical aberration–corrected high-resolution AC–HAADF–STEM images, (g) Illustration of the structure. (h) STEM–EDS element mapping images. Reprinted with permission from Ref. [79], Copyright 2022, American Chemical Society. | PMC9370661 | nanomaterials-12-02618-g006.jpg |
0.455875 | e5b3c8b0bb474e7daa30eb8253ab1458 | Mechanism of HER action on electrode surfaces in acidic media. Reprinted with permission from Ref. [85], Copyright 2019, American Chemical Society. | PMC9370661 | nanomaterials-12-02618-g007.jpg |
0.412705 | 8a86817d62c34bc18623c659bc842d4e | (a) Scheme demonstrating the synthesis process for depositing Pt nanoparticles on Co/NC nano-heterojunction materials, (b) TEM image, (c) HRTEM image, (d) HAADF--STEM image of a typical Pt4/Co sample, (e) LSV curves of the catalysts, (f) LSV curves for the Ptx/Co electrodes, (g) life tests of Pt4/Co and Pt/C electrodes, and (h) LSV curves of the Pt4/Co electrode before and after use for 24 h. All electrodes with the same catalyst loadings of 1 mg cm2 on carbon cloth (1 × 1 cm) were measured in Ar-saturated 0.5 M H2SO4 at a scan rate of 10 mV s−1. Reprinted with permission from Ref. [103], Copyright 2021, John Wiley and Sons. | PMC9370661 | nanomaterials-12-02618-g008.jpg |
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.