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
|
---|---|---|---|---|
0.466279 |
6de0f4c0020644b5b0be3ea29ec2a030
|
Receiver operating characteristic (ROC) curves for predictors of mortality in the COVID-19 cohort (n=74). CRP, C reactive protein; LDH, lactate dehydrogenase; gbCR, glass-bead test Clot Rate; gbPA, glass-bead test Peak Amplitude.
|
PMC9062462
|
bmjopen-2021-051971f04.jpg
|
0.394458 |
99b43645f4bb4d7693af857ff606281b
|
FE-SEM images of pLys-HAp composites: (a) α-pLys-HAp (20 mg), (b) α-pLys-HAp (30 mg), (c) α-pLys-HAp (40 mg), (d) ε-pLys-HAp (20 mg), (e) ε-pLys-HAp (30 mg) and (f) ε-pLys-HAp (40 mg). The insets in (a–f) show TEM images of each sample.
|
PMC9062467
|
c9ra01764j-f1.jpg
|
0.48906 |
f9baa4fd2fbf491e8095d2ac45ef2bbe
|
(A) X-ray diffraction patterns of (a) α-pLys-HAp (20 mg), (b) α-pLys-HAp (30 mg), (c) α-pLys-HAp (40 mg), (d) ε-pLys-HAp (20 mg), (e) ε-pLys-HAp (30 mg) and (f) ε-pLys-HAp (40 mg), respectively (JCPDS card no. 09-0432). (B) FT-IR spectra of (a) α-pLys-HAp (20 mg), (b) α-pLys-HAp (30 mg), (c) α-pLys-HAp (40 mg), (d) pure α-pLys, (e) ε-pLys-HAp (20 mg), (f) ε-pLys-HAp (30 mg), (g) ε-pLys-HAp (40 mg) and (h) pure ε-pLys, respectively.
|
PMC9062467
|
c9ra01764j-f2.jpg
|
0.392276 |
12caa311ae9043e9bb659cf8f0f37133
|
STEM images and EDX maps of elements of (A) α-pLys-HAp (40 mg) and (B) ε-pLys-HAp (40 mg). Yellow, green and red colours display nitrogen, calcium and phosphorus elements, respectively.
|
PMC9062467
|
c9ra01764j-f3.jpg
|
0.545375 |
9675ca4c68824910b2b8a1681861c5d8
|
Adsorption isotherm curves for GOX on α-pLys-HAp (40 mg) and ε-pLys-HAp (40 mg).
|
PMC9062467
|
c9ra01764j-f4.jpg
|
0.404961 |
ec17d88bd3c5458095b834ec4bf467b7
|
(A) Fluorescence spectra of free and immobilised GOX. (B) The three-dimensional structure of GOX was obtained from Protein Data Bank; ID: 1CF3. Hydrophobic amino residues in GOX were emphasised using red colour.
|
PMC9062467
|
c9ra01764j-f5.jpg
|
0.4578 |
a295e7a954294781a99ea11f1c3feb16
|
The Lineweaver–Burk plots of free and immobilised GOX.
|
PMC9062467
|
c9ra01764j-f6.jpg
|
0.521998 |
f7d7b7c09c944cb39558a199990cc267
|
Remaining activity of GOX immobilised on pLys-HAp in cycling test.
|
PMC9062467
|
c9ra01764j-f7.jpg
|
0.521671 |
57c45088931d4b0cb7bf71cb8c630f9d
|
The glucose-sensitivity using GOX (free and immobilised on pLys-HAp). The inset shows the plot in the range of 4–80 μM of glucose.
|
PMC9062467
|
c9ra01764j-f8.jpg
|
0.497781 |
2994f7add3fb4b53ab6a01a96092fd15
|
Cyclic voltammograms of (A) graphene/imogolite/GOX/α-pLys-HAp (40 mg)-modified GCE and (B) graphene/imogolite/GOX/ε-pLys-HAp (40 mg)-modified GCE for the addition of 0–2.0 mM glucose in O2-saturated Dulbecco's PBS (pH 7.3) at a scan rate of 100 mV s−1. The insets in (A) and (B) show the linear plots for the concentration of glucose vs. peak current.
|
PMC9062467
|
c9ra01764j-f9.jpg
|
0.46558 |
653b98f0a74a45feb399cc4e0d9b237f
|
Aneurysm wall pathophysiology and progression. (a) The normal arterial vessel wall consists of an inner endothelial layer (tunica intima) followed by the two strength layers: internal elastic lamina and the smooth-muscle of the tunica media. Dotted line represents the region of the cross-sectional image displayed on the left (b) Regions of high WSS (blue arrow) (e.g., vessel bifurcations and outer wall of curved vessels) trigger an inflammatory cascade, increased protease activity, and breakdown of the tunica media and internal elastic lamina. Loss of integrity of the strength layers of the vessel wall cause outpouching of the vessel wall and formation of an early aneurysm bulge. (c) Enlargement of the aneurysm sac alters the flow dynamics, leading to areas of low WSS and high WSS (blue arrows). Areas of low WSS can develop thrombus, which trigger further inflammation and wall breakdown. Areas of persistent high WSS continue to experience protease activity and inflammation leading to continues degradation of the tunica media and internal elastic lamina.
|
PMC9062958
|
SNI-13-182-g001.jpg
|
0.441664 |
d2abe0ec2a784d859ffccd61aa27e0ab
|
High-risk aneurysm morphologies. (a) Standard saccular aneurysm morphology. (b) Aneurysm with daughter sac. (c) Aneurysm with high aspect ratio. (d) Aneurysm with high size ratio.
|
PMC9062958
|
SNI-13-182-g002.jpg
|
0.407504 |
de3b70d59d234f59b607240f74e49018
|
The total length stacked bar chart of 26 Ligusticum plastomes composed of four regions (LSC, IRb, SSC, and IRa). The numbers on the bar represent the length of the four regions. A–K Represents the genes at IR/SC borders. A ycf2; B petB; C rpl22; D rps19; E ycf1/ndhF; F ycf1; G trnN-GUU-ndhF; H trnL-CAA-trnH-GUG; I petB-trnH-GUG; J rps19-trnH-GUG; K rpl2-trnH-GUG. All the SSC/IRa borders are ycf1, which is indicated by asterisks
|
PMC9063207
|
12862_2022_2010_Fig1_HTML.jpg
|
0.543402 |
fba8fa4f62e94f9dad939fbb6fb0a371
|
Comparison of the GC content (GC%) of 26 Ligusticum plastomes using a radar-plot. From inside to out: SSC GC%, LSC GC%, Total GC%, IR GC%, and rRNA GC%. The background colors of purple, green, blue, and pink represent Selineae, Sinodielsia Clade, Acronema Clade, and East-Asia Clade, respectively
|
PMC9063207
|
12862_2022_2010_Fig2_HTML.jpg
|
0.424184 |
4fe0cf129fc343e1b38612ab9747db57
|
The RSCU values of 53 merged protein-coding sequences for 26 Ligusticum plastomes. Color key: the red values indicate higher RSCU values and the blue values indicate lower RSCU values
|
PMC9063207
|
12862_2022_2010_Fig3_HTML.jpg
|
0.389253 |
e1276ce301a94e1d812f5cdc27247e10
|
The dN/dS (ω) and nucleotide diversity (Pi) of the 79 protein-coding sequences within 26 Ligusticum plastomes
|
PMC9063207
|
12862_2022_2010_Fig4_HTML.jpg
|
0.476195 |
85f8bb8e3b2c4802b29629ceba6905bb
|
Phylogenetic relationships inferred from Maximum likelihood (ML) and Bayesian inference (BI) analyses based on 66 complete plastomes within Apiaceae. The bootstrap support values (BS) and posterior probabilities (PP) are listed at each node
|
PMC9063207
|
12862_2022_2010_Fig5_HTML.jpg
|
0.442818 |
ecc365fec3294bf38b3bd7b5f8168a48
|
Phylogenetic relationships of 66 Apiaceae species inferred from 76 common protein-coding sequences based on the coalescent-based approach using ASTRAL. The local posterior probabilities (LPP) are listed at each node
|
PMC9063207
|
12862_2022_2010_Fig6_HTML.jpg
|
0.433789 |
69aa5f9afa5946dba1c812dda3a35124
|
UV-vis absorption spectra of P(O-DBND-2T), P(N,O-DBND-2T) and P(N-DBND-2T) in diluted o-DCB (a) and as the thin film (b).
|
PMC9063523
|
c9ra01545k-f1.jpg
|
0.405163 |
37422eea578c44829ecbb238f0e013b1
|
Photovoltaic characteristics: J–V curves (a) and EQE (b) plots of polymer : PC71BM (1 : 2 w/w) optimal solar cells under the illumination of AM 1.5 G 100 mW cm−2.
|
PMC9063523
|
c9ra01545k-f2.jpg
|
0.456152 |
422396f7326a4e389f81d2a9a9358176
|
GIWAXS curves of polymer neat films (a) and TEM images of active layers containing P(N,O-DBND-2T) : PC71BM (b), P(O-DBND-2T) : PC71BM (c) and P(N-DBND-2T) : PC71BM (d) in weight ratio of 1 : 2.
|
PMC9063523
|
c9ra01545k-f3.jpg
|
0.446769 |
694e511aa61545a28e2f9e95d46029d0
|
Chromatograms (distribution of n-alkanes for the test oils).
|
PMC9064330
|
c9ra02775k-f1.jpg
|
0.429916 |
56fdbc764ede45ecadaf029aab91d789
|
Submerged oils formed observed by UV epi-fluorescence. Fluorescing green indicates the oil droplets.
|
PMC9064330
|
c9ra02775k-f2.jpg
|
0.477304 |
40babe3e3bb043c8ac805c5ff7f3a513
|
Adsorption isotherms of the three test oils. Note: the dash lines are the fitted line using the Langmuir isotherm.
|
PMC9064330
|
c9ra02775k-f3.jpg
|
0.466554 |
3a3382365fc84070a5a7e2d46fce8f3b
|
The kinetics of submerged oil formation for the three test oils. Note: the dashed lines are fitted by eqn (4).
|
PMC9064330
|
c9ra02775k-f4.jpg
|
0.450702 |
994bdf1c6a484ccb9435a302c9ffe5c7
|
Modeling of predicting the submerged oils formation as a function of sediment concentrations. The dash lines represent the model results computed using eqn (7) based on Kd = 0.5 and 1.5 mL mg−1. Note: The use of chemical dispersant and change of salinity lead to deviate from model results.
|
PMC9064330
|
c9ra02775k-f5.jpg
|
0.445267 |
0de5b0dabbeb4056876821eb75124976
|
Cell viability of ATP and UTP-treated cells. TNBC MDA-MB 231, Hs 578t and MDA-MB 468 cell lines and non-tumorigenic immortal mammary epithelial MCF-10A cells were treated for 48 hours with increasing concentrations of ATP or UTP, and cell viability was measured with the PrestoBlue HS assay. Error bars represent standard deviations calculated from three independent experiments performed in triplicate. The student’s t-test was applied to the to ascertain significance. * represents p < 0.05 and ** represents p < 0.01 when comparing ATP to UTP.
|
PMC9065442
|
fonc-12-855032-g001.jpg
|
0.397869 |
4bd33f59abce4da0915850e75dff551b
|
Cell viability of ATP-treated cells in the presence of P2RX inhibitors. TNBC cell lines and MCF-10A cells were treated for 48 hours with increasing concentrations of ATP in the presence of the P2RX inhibitor Iso-PPADS (20 µmol/L), the P2RX7 inhibitor A438079 (20 µmol/L) or the P2RX4 inhibitor 5-BDBD (20 µmol/L) or vehicle addition, and cell viability was measured using the PrestoBlue HS assay. Error bars represent standard deviations calculated from three independent experiments performed in triplicate. One way ANOVA with Tukey’s HSD was applied to ascertain significance. * represents p < 0.05 and ** represents p < 0.01 when comparing vehicle addition to Iso-PPADS, A438079 or 5-BDBD.
|
PMC9065442
|
fonc-12-855032-g002.jpg
|
0.549365 |
0bc4a7c8d8524dfe821ac99ff76e3b30
|
Comparing eATP release from paclitaxel-treated cells in the presence of inhibitors or vehicle addition. (A) TNBC and MCF-10A cells were treated with increasing concentrations of paclitaxel and the nucleoside phosphohydrolase inhibitors POM-1 (E-NTPDase inhibitor, 10 µmol/L), PSB 069 (E-NTPDase inhibitor, 10 µmol/L), ENNP1 inhibitor C (ENPP1 inhibitor, 10 µmol/L) or vehicle addition for six hours, and cell viability was measured using the PrestoBlue HS assay. Standard deviation was calculated from three independent experiments performed in triplicate. (B) eATP concentrations were measured in the supernatants of TNBC and MCF-10A cells after six hours of treatment with increasing concentrations of paclitaxel and nucleoside phosphohydrolase inhibitors or vehicle addition. Standard deviation was calculated from three independent experiments performed in triplicate. One way ANOVA with Tukey’s HSD was applied to ascertain significance. * represents p < 0.05 and ** represents p < 0.01 when comparing vehicle addition to PSB 069. We highlighted just the significance in the presence of PSB 069 because the cell viability results were consistently significant.
|
PMC9065442
|
fonc-12-855032-g003.jpg
|
0.509328 |
820bbffae5e44d80b266e6253a38b387
|
Examining the influence of P2RX inhibitors in combination with E-NTPDase inhibitor on cell viability and eATP release in paclitaxel-treated cells. (A) Paclitaxel-treated breast cancer MDA-MB 468 cell lines were treated for six hours with P2RX7 inhibitor A438079 (20 µmol/L) or P2RX4 inhibitor 5-BDBD (20 µmol/L) in the presence or absence of PSB 069 (10 µmol/L), and cell viability was measured by applying PrestoBlue HS assay. Standard deviation was calculated from three independent experiments performed in triplicate. We used the same values for both graphs for vehicle addition and PSB 069. (B) eATP concentrations were measured in the supernatants of paclitaxel-treated MDA-MB 468 cells after six hours of treatment. Standard deviation was calculated from three independent experiments performed in triplicate. We used the to the same values for both graphs for vehicle addition and PSB 069. The student’s t-test was applied to the applicable assays to ascertain significance. * represents p < 0.05 and ** represents p < 0.01 for A438079 and PSB 069 or 5-BDBD and PSB 069 when compared to PSB 069 alone.
|
PMC9065442
|
fonc-12-855032-g004.jpg
|
0.440819 |
f1044067a9764cc4b70cf2127f531256
|
Determining relative eATP content and cell viability in paclitaxel-treated cells in the presence of ivermectin or vehicle addition. (A) The graphs represent cell viability as measured using the Presto Blue HS assay +/- standard deviation from three independent experiments performed in triplicate in TNBC and MCF-10A cells after six hours of treatment with increasing concentrations of paclitaxel and the P2RX4 and P2RX7 activator ivermectin (10 µmol/L) or vehicle addition. (B) eATP content was measured in the supernatants of paclitaxel-treated TNBC and MCF-10A cell lines in the presence of the P2RX4 and P2RX7 activator ivermectin (10 µmol/L) or vehicle addition. The student’s t-test was applied to the applicable assays to ascertain significance. * represents p < 0.05 and ** represents p < 0.01 when comparing ivermectin to vehicle addition.
|
PMC9065442
|
fonc-12-855032-g005.jpg
|
0.428031 |
9426325427304cc4b3b60e348472132d
|
mRNA and protein expression analysis of P2RX4 and P2RX7 for all cell lines. (A) qRT-PCR was performed on mRNA of TNBC cell lines and MCF-10A cells using specific primers for P2RX4 and P2RX7. * represents p < 0.05 and ** represents p < 0.01. TNBC cell lines, MCF-10A cells, and HEK 293T cells transfected with P2RX4 or P2RX7 as positive controls were probed for (B) P2RX4 and (C) P2RX7, and GADPH was used as a loading control for western blot analysis repeated twice. HEK 293T cells transfected with P2RX7 were loaded at increasing protein concentrations of 1.0 µg, 2.5 µg, and 5.0 µg combined with lysates of control vector-transfected cells to keep the total loaded protein the same in each lane. Densitometry analysis was performed using Image Studio on the 75 kDa P2RX7 band. The student’s t-test was applied to the applicable assays to ascertain significance. * represents p<0.05 and ** represents p < 0.01 relative to MCF-10A; + represents p < 0.05 and ++ represents p < 0.01 relative to HEK293-empty vector transfected. (D) The calculated difference in mean fluorescence intensity (MFI) values between TNBC cell lines, MCF-10A cells, and HEK 293T cells transfected with P2RX4 or P2RX7 as positive controls stained with P2RX4 or P2RX7 specific antibody and the isotype control for the different cell lines examined. * represents p < 0.05 and ** represents p < 0.01 relative to MFI difference in MCF-10A cells; + represents p < 0.05 and ++ represents p < 0.01 relative to MFI difference in HEK293-empty vector transfected. O/E represents overexpressed.
|
PMC9065442
|
fonc-12-855032-g006.jpg
|
0.525798 |
c7accb745a4e4ff29fab1284f56a9a3c
|
Schematic displaying our proposed model for ATP release. Our proposed model suggests that ivermectin activates P2RX4 and P2RX7 leading to the release of ATP and the more ATP that accumulates extracellular can promote cell death especially in the presence of paclitaxel. In addition, the breakdown of ATP can be prevented in the presence of E-NTPDase inhibitors POM-1 or PSB 069. However, the release of ATP can be prevented in the presence of P2RX4 inhibitors 5-BDBD or Iso-PPADS or P2RX7 inhibitors A438079 or Iso-PPADS.
|
PMC9065442
|
fonc-12-855032-g007.jpg
|
0.411999 |
6c4b0d2dbf51406d833299e7969b9eb1
|
XRD analysis of vanadium-bearing titanomagnetite concentrates.
|
PMC9065754
|
c9ra03271a-f1.jpg
|
0.42295 |
229139b078274d83b64db7e623257c2d
|
Worn surface and wear debris of sintered samples. (a) and (b) without rare-earth oxide, (c) (d) La-0.2 wt%, and (e) and (f) Ce-0.4.
|
PMC9065754
|
c9ra03271a-f10.jpg
|
0.520469 |
a80b5c9dc0ad4068a1f8eb8088c7519e
|
XPS spectrum of the worn surface of sintered sample with different rare-earth addition: (a) carbon, (b) Fe on the worn surface of Re-0, (c) Fe on the worn surface of La-0.2, (d) Fe on the worn surface of Ce-0.4.
|
PMC9065754
|
c9ra03271a-f11.jpg
|
0.376181 |
aa44216db6744fa1890c89def179a580
|
XRD analysis of the pre-reduced powders.
|
PMC9065754
|
c9ra03271a-f2.jpg
|
0.577574 |
ae521a5b8b8f406fb8aa3e2b8b4b949f
|
Diagram of block-on ring tester (a) grinding wheel, (b) specimen, and (c) load.
|
PMC9065754
|
c9ra03271a-f3.jpg
|
0.438372 |
354d73a128b14454a2a0b1ad794f22bb
|
Microstructure of sintered sample (a) and EDS analysis of area A (b), B (c) and D (d).
|
PMC9065754
|
c9ra03271a-f4.jpg
|
0.36366 |
989af5fd0a7042c789bdfd1d42da03ee
|
XRD pattern of the sintered sample without rare-earth oxide.
|
PMC9065754
|
c9ra03271a-f5.jpg
|
0.431699 |
1a0e538021de42e4ab939800f010da8b
|
SEM micrographs of powders (a) without and with (b) 0.2 wt% La2O3, (c) 0.4 wt% La2O3, (d) 0.4 wt% CeO2 and (e) 0.6 wt% CeO2.
|
PMC9065754
|
c9ra03271a-f6.jpg
|
0.404877 |
38e19f5e8bea40c0a18d33029cee053d
|
Microstructure of the sintered specimens with (a) 0.2 wt% La2O3, (b) 0.2 wt% CeO2, (c) 0.4 wt% La2O3, (d) 0.4 wt% CeO2, (e) 0.6 wt% La2O3, (f) 0.6 wt% CeO2.
|
PMC9065754
|
c9ra03271a-f7.jpg
|
0.412863 |
ddc6af94931c4d14a3c1d310cf26407f
|
Hardness and relative density of Fe-based friction material with different rare earth.
|
PMC9065754
|
c9ra03271a-f8.jpg
|
0.402829 |
eb4a9abe0550420dbd42bc37d6315e1e
|
Wear rate and friction coefficient of Fe-based friction material with different rare earth addition.
|
PMC9065754
|
c9ra03271a-f9.jpg
|
0.420233 |
9e80b30cdb2d4630ab8e77a2f734ae01
|
Identification of cofilin-rich regions at the base of neuronal growth cone filopodia.a–c Immunofluorescence images of phalloidin (Alexa Fluor 488) alone (a), cofilin (Alexa Fluor 594) alone (b), and a merge of the two (c). Most of the cofilin signal emanates from linear aggregates near the base of filopodia. The white dashed line in c marks the position of the lamellipodial veil. d Merged immunofluorescence image of a growth cone with fascin (Alexa Fluor 488-green) and cofilin (Alexa Fluor 594-red) labeled, showing cofilin-rich regions at the base of filopodia as in c. Bottom-left inset: Split view of the boxed-out region. The white arrow points to the same location in each image and shows the point at which the fascin signal drops off and the cofilin signal intensifies. e Representative line scan intensity profile of a single filopodium showing the distribution of fascin and cofilin. The image above the graph shows a close-up view of the measured filopodium and the location where the line profile was drawn. The transition region is marked by two dashed lines. The images in a–c are representative images, but two independent experiments showed similar localization of cofilin and actin. d is a representative image from one of three independent experiments. Scale bars: (c) 5 µm (this also corresponds to a and b), d 5 µm. Source data are provided as a Source Data file.
|
PMC9068697
|
41467_2022_30116_Fig1_HTML.jpg
|
0.416944 |
b29faa2b41624e7b913720228450684d
|
Structural features of neuronal growth cone filopodia and their associated cofilactin bundles.a Overview image of a cryo-preserved growth cone on a Quantifoil EM grid. The green and red boxes represent growth cone regions similar to where the tomograms in b and c were imaged, respectively. b, c 5 nm-thick slices of tomograms from the tip (b) and base (c) of growth cone filopodia. In b, a bundle of actin filaments fills the entire cytoplasm. In c, branched networks of individual actin filaments can be seen surrounding a central bundle of hyper-twisted cofilactin filaments. White arrows point to the bundle. Lower-left insets: 68 nm-thick transverse cross-sections through each bundle, illustrating the hexagonal packing of filaments. The blue line in the main images show the plane from which the insets were taken. Bottom insets: Subtomogram averages of filament pairs in filopodial tips (below b) or in cofilactin bundles at the filopodial bases (below c). Cofilactin filaments have a shorter helical twist than F-actin and are out of phase with adjacent filaments. d EM map (blue) resulting from the subtomogram averaging of actin filaments in filopodial tips, and rigid body fitting of a previously reported atomic structure for F-actin (PDB ID: 6T1Y; green). e EM map (blue) resulting from the subtomogram averaging of cofilactin filaments near the base of filopodia, and rigid body fitting of a previously reported atomic structure of cofilactin (PDB ID: 3J0S; actin is green and cofilin is red). f Segmented filopodial protrusion with a schematic of filament centerlines overlaid (red). These lines are comprised of a series of points that were used for nearest neighbor analysis. g Nearest neighbor histograms showing the cumulative total of three normal actin filopodial bundles (green) and three cofilactin bundles (red). Scale bars: (a) 5 µm, (b) 200 nm (this also corresponds to the image in c), f 100 nm.
|
PMC9068697
|
41467_2022_30116_Fig2_HTML.jpg
|
0.459147 |
fa5d3a1a29b444e9b8f99f456af67748
|
Higher-order structure of filopodial bundle types.a Schematic model of the higher-order structure of actin (fascin-linked) and cofilactin bundles in growth cone filopodia (as they appear looking down the long axis of the bundle). Filaments in both bundle types are organized in layers and hexagonally packed (blue hexagons). Filaments in the actin bundle are all twisting in phase with one another, but in the cofilactin bundle filaments are rotated 90° with respect to their neighbors in the same layer. This creates columns of filaments that are oriented similarly to one another. The yellow dots on filaments in the hexagon correspond to filaments in b. b 17 nm-thick slices through tomograms of a filopodial tip (top) and a cofilactin bundle (bottom). In the cofilactin bundle, brackets show the wide portion of the helical twist while red arrows show the thin portion. The yellow dots correspond to the dots in a. For instance, the bracketed filament in layer two is directly between the two filaments in layers 1 and 3, only on a different Z-plane. Scale bars in b are 20 nm.
|
PMC9068697
|
41467_2022_30116_Fig3_HTML.jpg
|
0.4784 |
d256bdc063b14fdebca310516b361aad
|
Potential conflict between fascin and cofilin, and potential cofilactin interactions.a, b Top-down views of fascin-linked actin (a) and cofilactin (b) filaments. Both pairs of filaments are shown at interfaces where they are in phase with one another. In a, fascin (blue) is shown in its actin-binding pockets (22), two sites that are sterically blocked by the presence of cofilin in our model of cofilactin bundles (b). F-actin is green, and cofilin monomers are red. c Sideview of model cofilactin hexagonal unit. Dashed regions labeled 1 and 2 illustrate cross-sections shown in d and e, respectively. F-actin is green and cofilin is red. d, e Cross-sections through different regions of the hexagonal unit shown in c. Arrows point to the place where cofilin monomers on neighboring filaments are closest to each other (~2 nm apart at their closest residues). f Zoomed-in view of neighboring cofilins like those in the boxed-out region in e. Cys39 and Cys147 are displayed on each (cyan).
|
PMC9068697
|
41467_2022_30116_Fig4_HTML.jpg
|
0.52466 |
6e5b30cf550548de9dd1db082cdc4531
|
Tomography of cofilactin within filopodia.In all images, red arrows signify cofilactin and green arrows show normal F-actin. The top-right corner of each image shows their distance from the lamellipodial veil (positive values are distal from the veil and negative values are proximal from it). a A 13.3 nm-thick tomographic slice where individual cofilactin filaments are scattered throughout a fascin-linked actin bundle. b 13.3 nm-thick tomographic slice through a prospective transition region where a clear boundary (represented by the dashed line) exists between the F-actin on top and the cofilactin on bottom. c 13.3 nm-thick tomographic slice showing a pure cofilactin bundle). Scale bars: 50 nm.
|
PMC9068697
|
41467_2022_30116_Fig5_HTML.jpg
|
0.447194 |
c7ff18d3280148aeabf6c95a8d27753c
|
Cofilactin bundles facilitate bending and breaking of filopodial protrusions.a Single image of a whole-cell expressing tdTomato-Lifeact (pseudocolored green) and EGFP-cofilin (pseudocolored red). b, c Maximum intensity projections (MIP) of lower and upper boxed regions in a, respectively. MIPs include 40 images at 3 s intervals for 2 min total. Filopodia in b are examples of “resting” filopodia and (c) shows a “searching” filopodium. Dashed arrows indicate the direction of actin retrograde flow, and arrowheads indicate inflection points seen along the flexing filopodial bundles. The white dashed line marks the position of the lamellipodial veil. d–f 2-min movie montages showing behaviors exhibited by the cofilin-rich filopodia designated with the corresponding label in a. Arrowheads follow either a kink/breaking point (d), a cofilactin bundle “wave” (e), or an inflection point in a bending filopodium (f). In e, the wave is caused by the filopodial tip moving from left to right, which drags the attached base behind it in a flexible, wave-like motion. The localization of EGFP-cofilin and tdTomato-Lifeact shown in a was replicated in multiple cells from two independent experiments, and similar results were also seen using other fluorescent protein combinations (Supplementary Fig. 5). Scale bars: (a, b) 5 µm (scale bar in b also corresponds to image in c), (d) 500 nm (also corresponds to (e) and (f)).
|
PMC9068697
|
41467_2022_30116_Fig6_HTML.jpg
|
0.434371 |
1478987d4ff34e8281fb9ebb3de9d68d
|
Tomograms of growth cone filopodia in different dynamic states.In all images, red arrows signify cofilactin and green arrows show normal F-actin. a 13.3 nm-thick tomographic slice showing a ~90° kink in a growth cone filopodium. This filopodium is likely in the process of severing, like in Fig. 6d and Supplementary Movie 7. Cofilactin filaments can be seen near the apex of the kink in the bundle. b 13.3 nm-thick tomographic slice where the proximal portion of a filopodium appears to be moving from right to left in a wave-like motion. This is similar to the motion exhibited by the filopodium shown in Fig. 6e and Supplementary Movie 8. The dashed boxes on top and bottom represent the regions shown in the top and bottom zoomed-in insets (white boxes), respectively. The top zoomed-in view shows a cofilactin filament forming an S-curve through the fascin cross-linked bundle. The cofilactin filament is moving through the Z-axis, so the middle of the S-curve is not visible in this image. The bottom zoomed-in view shows, similar to the kink from (a), cofilactin filaments near the crest of the wave. c 13.3 nm-thick tomographic slice of a bend in a distal filopodial region. Here, cofilactin and actin coexist as separate bundles wrapping around one another. This tomogram resembles that shown in the movie from Fig. 6f and Supplementary Movie 9. Inset: TEM overview image showing the location of the tomogram in the main panel. Scale bars: a, b, and c represent 100 nm.
|
PMC9068697
|
41467_2022_30116_Fig7_HTML.jpg
|
0.42792 |
33b083be14ba4143b3b8d84d12396ab4
|
Current model.Schematic of filopodial actin bundle in a rigid (a) and flexible (b) state. As cofilactin filaments permeate the fascin-linked region of the filament, they competitively dislodge fascin cross-linkers and increase the flexibility of filopodial bundles. In the proximal region of the bundles, cofilactin filaments prevail and are cross-linked through either cofilin oligomers/self-association or by some, as of yet, unknown cross-linker.
|
PMC9068697
|
41467_2022_30116_Fig8_HTML.jpg
|
0.420134 |
9b02e2d755f449399fd10eb3b00db19e
|
Histidine levels were significantly higher in the hypersensitivity group. (A) Heatmap and cluster analysis of significant differential metabolites between the normal group (group 1) and the hypersensitivity group (group 2). (B) Z-score graph of significant differential metabolites. l-Histidine levels in the normal group were significantly higher than those in the hypersensitivity group. (C–F) Four significant differential metabolites levels between the normal and hypersensitivity groups included l-histidine, urocanic acid, myristicin, and d-aldose. (G) Pathway enrichment analysis. Histidine metabolism showed a high impact on hypersensitivity (*p < 0.05, **p < 0.01).
|
PMC9068896
|
fphar-13-827446-g001.jpg
|
0.389697 |
9e6c0d34e45e457a8d1b4b30ff31c223
|
Receiver operating characteristic (ROC) curves and content level of 30 differential metabolites.
|
PMC9068896
|
fphar-13-827446-g002.jpg
|
0.436613 |
ef857c786e5b48d6bc96b1a00226a019
|
Histidine supplement-enhanced PLD-induced HSR in rat model. (A) H&E staining (100x). PLD injection could induce pulmonary edema. Histidine supplement-aggravated pulmonary edema, but this phenomenon could be alleviated by histidine decarboxylase inhibitor: bromo-3-hydroxybenzoic acid (BHBA). (B) Toluidine blue staining (mast cell staining) (200x). Histidine could increase mastocyte infiltration in the trachea and lungs and significantly increase mastocyte degranulation in the trachea and larynx after PLD injection. (C) Degranulation and undegranulation mast cell counting. (D) Histidine supplement group showed significantly increased IgE levels. BHBA treatment could decrease IgE levels significantly (PLD, PEGylated liposomal doxorubicin; HSR, hypersensitivity reaction; *p < 0.05, **p < 0.01, ***p < 0.001).
|
PMC9068896
|
fphar-13-827446-g003.jpg
|
0.492023 |
d2252878f3ca49a3bfd0da03ba1b8dcf
|
Corona virus binds to Lung Epithelial Cells which causes the activation of inflammatory cells and cytokine storms
|
PMC9069147
|
IJPVM-13-45-g001.jpg
|
0.503843 |
a01b8204cad047d882e988124f40658c
|
Schematic representation of the multimodal X-ray microscopy experimental setup. KB mirrors focus X-rays down to a focal spot of about 70 nm and with a 1.2 mrad divergence. The sample is mounted on a motor stack enabling xyz translation and θ, φ rotation, i.e. around the x and y axes, respectively. The sample can be translated in and out of the X-ray beam focus via translation along z (total travel range 46 mm), effectively tuning the X-ray beam size. An ion chamber is positioned downstream of the KB chamber to monitor the incoming X-ray photon flux. An optical in-line microscope provides a view onto the sample along the optical z axis. A fluorescence detector is positioned 15 mm away from focus, with a 15° orientation with respect to the focal plane. An in-line area detector is positioned 1.12 m downstream of the focus to collect holograms.
|
PMC9070709
|
s-29-00807-fig1.jpg
|
0.490743 |
991d191b7afe4bb083da3c076e2b4174
|
Preliminary reconstructions from a 5 × 5 holography scan covering a potential ROI of a section of an Os-stained human peripheral sural nerve biopsy from a healthy male. These qualitative results were used to delimit a ROI on which to perform a nanoscale X-ray fluorescence scan, as more conveniently annotated in Fig. 3 ▸. Further processing followed. The frame number is annotated on each frame, revealing the scanning sequence.
|
PMC9070709
|
s-29-00807-fig2.jpg
|
0.476043 |
e3222abd6f904fe1ab842ceb8a0d4d6a
|
Quantitative result obtained from processing the images from Fig. 2 ▸. The ROI on which a nanoscale X-ray fluorescence scan was performed is highlighted by the white rectangle. A blood vessel (blue arrow) as well as myelin layers surrounding several axons (red arrows) can be identified. The intensity scale represents relative electron density variations and for convenience was converted into units of Å−3.
|
PMC9070709
|
s-29-00807-fig3.jpg
|
0.496383 |
9b0bff561368482e9b4479a1af7d7b96
|
(a)–(c) Area mass density maps obtained via nanoscale X-ray fluorescence emission spectroscopy for (a) Cl, (b) Fe and (c) Os within a ROI of a section of an Os-stained human peripheral sural nerve biopsy from a healthy male. Linear intensity scales are shown in units of ng mm−2. (d) RGB representation of the same region including electron density information from the holography scan: red, green and blue channels represent electron density, Fe and Os area mass densities, respectively. A 10 µm scale bar for all images is given in (a).
|
PMC9070709
|
s-29-00807-fig4.jpg
|
0.489397 |
55d3bd280c2b4912bd0bc4ea90bb5d21
|
Postprandial 1,2-dicarbonyl compound concentrations in older (25 months) and younger (5 months) male mice. Data are shown as Violin plots; dotted lines represent median and quartiles. Student’s t-test for normally distributed data and Mann–Whitney U test for nonnormally distributed data were used to assess differences between age groups. 25 months old (n = 11), 5 months old (n = 14). 3-DG = 3-deoxyglucosone; GO = glyoxal; MGO = methylglyoxal; 25 M = 25 months old; 5 M = 5 months old; p < .05.
|
PMC9071428
|
glab331_fig1.jpg
|
0.456125 |
d58d26f7c0a44d7eaa2da5280acedc4b
|
Fasting 1,2-dicarbonyl compound concentrations in older and younger women (A) and men (B). Data are shown as Violin plots; dotted lines represent median and quartiles. Student’s t-test for normally distributed data and Mann–Whitney U test for nonnormally distributed data were used to assess differences between age groups. YW (n = 19), OW (n = 19), YM (n = 15), OM (n = 15). 3-DG = 3-deoxyglucosone; GO = glyoxal; MGO = methylglyoxal; YW = younger women; OW = older women; YM = younger men; OM = older men; p < .05.
|
PMC9071428
|
glab331_fig2.jpg
|
0.476952 |
62f927fad4ab49faabb5e8d6b86b39f1
|
Postprandial 1,2-dicarbonyl concentrations and response (iAUC) to a dextrose challenge in older and younger women and men. Data are shown as mean ± SD. Repeated measures analysis of variance was used to examine changes over time and differences between age groups, and Mann–Whitney U test was used to assess age differences of the postprandial response (iAUC). *Significantly different between age groups. YW (n = 19), OW (n = 19), YM (n = 15), OM (n = 15). iAUC = incremental area under the curve; 3-DG = 3-deoxyglucosone; GO = glyoxal; MGO = methylglyoxal; YW = younger women; OW = older women; YM = younger men; OM = older men; p < .05.
|
PMC9071428
|
glab331_fig3.jpg
|
0.405949 |
61457424ace84c04a9025d6761df8c9c
|
Associations of glucose response with 3-DG (A) and GO (B) response as well as insulin response with MGO response (C) to a dextrose challenge. (A) No significant correlation in women, men: r = 0.782, p = .008. (B) No significant correlation in women, men: r = 0.707, p = .022. (C) No significant correlation in women, men: r = 0.784, p = .007. YW (n = 19), OW (n = 19), YM (n = 15), OM (n = 15). iAUC = incremental area under the curve; 3-DG = 3-deoxyglucosone; GO = glyoxal; MGO = methylglyoxal; YW = younger women; OW = older women; YM = younger men; OM = older men; p < .05.
|
PMC9071428
|
glab331_fig4.jpg
|
0.457334 |
fe5311e4dd2c4323a6754a6d5e03aae0
|
Flow diagram of the simulation, benchmarking, and experimental data evaluation strategy presented in the manuscript. Briefly, SplattDR was developed to simulate dose–response scRNAseq data and validated based on experimental dose–response data. Simulated datasets were generated varying diverse parameters 10 times and then used to assess the performance of each test method. Each test method was also assessed using experimental data from the hepatic snRNAseq dose response dataset obtained from male mice gavage every 4 days for 28 days with 0.01, 0.03, 0.1, 0.3, 1, 3, 10 or 30 μg/kg TCDD. Related figures for each analysis from the main body are noted.
|
PMC9071439
|
gkac019fig1.jpg
|
0.42975 |
52a53f4cf5254c458e0a1d49a611aca5
|
Comparison of simulated and real dose–response data. (A) Relationship between gene-wise mean expression and percent zeroes for simulated and real dose–response data. Simulation data consisted of 10 000 genes and nine dose groups based on parameters derived from experimental dose–response snRNAseq data. Black line represents a fitted model to the experimental data from which the normalized root mean square deviation (NRMSD) of simulated data was determined. (B) Relationship between gene-wise mean expression and variance for simulated and experimental data. NMRSD was calculated for simulated data from the fitted model represented as a black line. (C) Distribution of log(fold-changes) in experimental and simulated data showing the median and minimum and maximum values. (D) Principal components analysis of simulated data colored according to simulated dose groups. (E) NMRSD estimated relative to fitted model in A,B for simulated data generated from initial parameters derived from published hepatic scRNAseq (two dose; GSE148339), hepatic whole cell (whole cell; GSE129516), and peripheral blood mononuclear cell (PBMC; GSE108313) datasets. (F) NMRSD estimated relative to model fitted to cell-type specific experimental dose–response data when simulated from initial parameters estimated from that same cell type. Box and whisker plots show median NMRSD, 25th and 75th percentiles, and minimum and maximum values.
|
PMC9071439
|
gkac019fig2.jpg
|
0.40547 |
9393420bace24831bf038a4622e7fc70
|
Classification performed of DE analysis tests. (A) ROCs estimated from simulated dose–response scRNAseq data for nine DE test methods including all genes expressed in at least one cell (unfiltered). (B) ROCs for nine DE test methods after filtering simulated dose–response scRNAseq data for genes expressed in only ≥5% of cells (low levels) in at least one dose group. (C) Precision-recall curves (PRCs) for nine DE test methods on unfiltered simulated dose–response scRNAseq data. (D) PRCs for nine DE test methods on filtered simulated dose–response scRNAseq data. Lines represent the mean values and shaded region reflects the standard deviation for 10 independent simulations. (E) Precision of DE test methods. (F) FPR of DE test methods. (G) MCC for test methods. (E–G) Box and whisker plots median values, 25th and 75th percentiles, and minimum and maximum values for 10 independent simulations. Points reflects values for each independent simulation. Panels display comparisons of unfiltered and filtered datasets.
|
PMC9071439
|
gkac019fig3.jpg
|
0.433991 |
07e550188fbf4a399430f9c372ce5640
|
Evaluation of Type I and II error control. (A) False positive rate (FPR) of 9 differential expression test methods estimated from negative control (0% DE genes) simulated dose–response scRNAseq data including all genes expressed in at least 1 cell (unfiltered) and genes expressed in only ≥ 5% of cells in at least one dose group (filtered). (B, C) Logistic regression models were fitted to negative control data to predict the probability of false positive identification using percent zeroes and mean expression as covariates. Lines represent the predicted probability of false positive classification with the shaded region representing the 95% confidence interval. (D) False negative rate (FNR) of nine differential expression test methods estimated from positive control (100% DE genes) simulated dose–response scRNAseq data including unfiltered and filtered datasets. (E, F) Logistic regression models were fit to positive control data. Lines represent predicted probability of false negative classification with shaded region representing the 95% confidence interval.
|
PMC9071439
|
gkac019fig4.jpg
|
0.496533 |
df576553fcc34d4585990e8a9d5f80c9
|
Matthews correlation coefficient (MCC) from sensitivity analyses of differential expression test methods. (A) MCC for nine DGEA test methods determined from simulated dose response data with varying number of cells per dose group. Simulations consisted of 5,000 genes with a probability of differential expression of 10% and 9 dose groups. (B) MCC for simulated data varying the cells numbers by dose group. The number of cells in each of the nine doses groups is shown on the right. (C) MCC for varying proportion of differentially expressed genes. (D) MCC when varying the mean fold-change (location) of repressed differentially expressed genes. (E) MCC for varying distribution of fold-change (scale) of differentially expressed genes. (F) MCC for varying dropout rates calculated as in Supplementary Table S3. Points represent median and error bars represent minimum to maximum values. Boxplots represent median, 25th to 75th percentile, and minimum to maximum values. Each analysis consisted of 10 replicate datasets including all genes expressed in at least one cell (unfiltered) and genes expressed in ≥5% of cells in at least one dose group (filtered).
|
PMC9071439
|
gkac019fig5.jpg
|
0.440178 |
3773fa32a7084a4e892b195a546cd6bc
|
Agreement of differential expression test methods on experimental dose–response data. (A) Upset plot showing the intersection size of genes identified as differentially expressed by nine different test methods in hepatocytes from the portal region of the liver lobule. (B) Intersect of differentially expressed genes in portal fibroblasts. (C) Intersect size in hepatic stellate cells. Vertical bars represent the intersect size for test methods denoted by a black dot. Horizontal bars show the total number of differentially expressed genes identified within each test (set sizes). Only intersects for which genes were identified are shown. Genes were considered differentially expressed when (i) expressed in >5% of cells within any given dose group and (ii) exhibit a |fold-change| ≥ 1.5. A heatmap in the upper left corner of each panel shows the pairwise AUCC comparisons for the 500 lowest P-values. (D) Relative proportion of cell types identified in each dose group of the real dataset for the cell types in (A–C). Experimental snRNAseq data was obtained from male mice gavaged with sesame oil vehicle (vehicle control) or 0.01–30 μg/kg TCDD every 4 days for 28 days. (E) Graph metrics for gene set enrichment analysis of portal fibroblasts grouped by similarity in gene membership. Violin plots show distribution of node-wise values for each test method. (F) Network visualization of significantly enriched (adjusted P-value ≤ 0.05) gene sets using the Bayes factor ranked genes of portal fibroblasts. Groups of ≥2 nodes were manually annotated following commonality in the gene set names. Each node represents a gene set with the size of the node representing the number of genes in a gene set, and edges connect nodes with ≥50% overlap.
|
PMC9071439
|
gkac019fig6.jpg
|
0.465634 |
992c09a99bbc4a58978ca6436e9a339f
|
Median ranking of differential expression test methods across all simulations. The median rank of each test method was calculated for AUPRC, AUROC, MCC, FNR and FPR. Tests were grouped according to intended application including fit-for-purpose tests developed for the analysis of dose–response datasets, multiple group tests, and two group tests. The overall rank represents the median value for the five key metrics presented here.
|
PMC9071439
|
gkac019fig7.jpg
|
0.393916 |
2ab64ca0949a48fa89c4d64172480268
|
The pattern classification of tumor microenvironment (TME). (a) Optimal number of clusters: K value calculated by the elbow method and gap statics algorithm. The ordinate axis represents total within sum of square; the abscissa axis represents the number of clusters K. (b) Consensus matrix heat map (K = 3): ConsensusClusterPlus was used for unsupervised class discovery (1000 iterations, k = 1 : 10). The optimal k value of 3 was determined using the elbow method and gap statics, combined with the correlation between the final classification and survival. (c) The distribution ratio of all kinds of immune cells in different TME clusters. (d) Clustering heat map of the distribution ratio of all kinds of immune cells in different TME clusters. (e) Survival analysis for different TME clusters: the red curve represents the TME cluster 1, the blue curve represents the TME cluster 2, and the yellow curve represents the TME cluster 3. The ordinate axis represents the probability of survival, and the abscissa axis represents the survival days.
|
PMC9071896
|
BMRI2022-5673810.001.jpg
|
0.456703 |
32d746738c974c51a9c41fcb1716fa75
|
Evaluation of TMEscore model and analysis of its correlation with mutation load. (a) The meta-analysis results of TMEscore model in different datasets: training set, testing set, TCGA, TCGA different stages, metastatic breast cancer, and prognosis evaluation of drug treatment. (b) Correlation analysis between TMEscore and mutation loads in different subtypes (basal, Her 2, Lum A, Lum B, and normal): the ordinate axis represents total mutations, and the abscissa axis represents TMEscore. (c) TMEscore box plots in different subtypes (basal, Her 2, Lum A, and Lum B). (d) Survival analysis results in four different subtypes of breast cancer (basal, Her 2, Lum A, and Lum B): the ordinate axis represents the probability of survival, and the abscissa axis represents the survival days. Different colors represent different subtypes. (e) Survival analysis of luminal A subtype after grouping according to TMEscore: the ordinate axis represents the probability of survival, and the abscissa axis represents the survival days. Different colors represent different TMEscore subgroups (high TMEscore and low). (f) Survival analysis of luminal B subtype after grouping according to TMEscore: the ordinate axis represents the probability of survival, and the abscissa axis represents the survival days. Different colors represent different TMEscore subgroups (high TMEscore and low).
|
PMC9071896
|
BMRI2022-5673810.002.jpg
|
0.425939 |
d27306a1ab2c43d4b12a7ecc0a8f1e47
|
Overview of mutations in 1039 TCGA-BRCA samples. (a) Tumor mutation profile. A1: variant classification of 1039 tumor samples. Missense mutations were the main mutation type in BRCA. A2: variant type of 1039 tumor samples. The source of mutations was mainly SNPs (mostly C>T) followed by indels. A3: SNV class of 1039 tumor samples. A4: variants per sample among 1039 tumor samples. A5: variant classification summary of 1039 tumor samples. A6: top 10 mutated genes in 1039 tumor samples. MUC4 was the most common mutated gene, followed by TTN. (b) Gene mutation distribution and phenotype in different TMEscore groups. B1: the distribution of mutations and mutation annotations of 24 genes in TMEscore high group. B2: the distribution of mutations and mutation annotations of 24 genes in TMEscore low group. (c) The frequency distribution of common gene mutations. Among them, the mutation rates of PIK3CA, TP53, KMT2C, GATA3, and MUC4 in the two subgroups have statistically significant differences.
|
PMC9071896
|
BMRI2022-5673810.003.jpg
|
0.451255 |
97b6d41d020c4b26be46b9c3c0d7938c
|
The analysis results of CNV. (a) The occurrence of chromosome arm-level amplification and deletion in different TMEscore groups: the abscissa axis represents the chromosome locus, and the ordinate axis represents the frequency of copy number alterations. Red represents the high TMEscore group, and the other one represents the low TMEscore group. ∗ represents statistical differences in frequency between the two groups. (b) Distribution of copy number amplification and deletion regions in high TMEscore group: 11q13.3 was the most significant in the amplified region, and 11q23.1 was the most significant in the deletion region. (c) Distribution of copy number amplification and deletion regions in TMEscore_low group: the most significant amplification region was located at 8q24.21, and the most significant deletion region was located at 8p23.2. (d) Ploidy analysis results in high and low TMEscore groups: ∗ represents statistical differences in frequency between the two groups. (e) Purity analysis results in high and low TMEscore groups: ∗∗ represents statistical differences in frequency between the two groups.
|
PMC9071896
|
BMRI2022-5673810.004.jpg
|
0.438062 |
bdbea8658c344e9f9cb116970067413e
|
The analysis results of miRNA and mRNA. (a) Functional annotation of differentially expressed miRNA: the ordinate axis represents different miRNAs, and the abscissa axis represents the functional annotations of miRNAs. (b) Volcano map of differentially expressed genes (DEG): the red part on the right shows the upregulated TOP 9 genes, and the blue on the left shows the downregulated top 2 genes. (c) Heat map of differentially expressed genes (DEG). (d) GO enrichment analysis of differentially expressed genes (DEG): the abscissa axis represents the number of genes, and the ordinate axis represents the results of CC (cellular component), BP (biological process), and MF (molecular function). (e) KEGG enrichment analysis of differentially expressed genes (DEG): the abscissa axis represents the number of genes, and the ordinate axis represents the corresponding pathways.
|
PMC9071896
|
BMRI2022-5673810.005.jpg
|
0.427068 |
1e225930a169458b96183c12b7784ab0
|
Comprehensive analysis results of tumor samples. (a) Volcano map of differentially methylated sites: 217 significantly different methylation sites were detected, and 67 methylation sites related to survival were obtained. (b) The survival analysis results of hsa-mic-1307 in different subgroups: the abscissa axis represents the survival time, and the ordinate axis represents the survival probability. The red curve represents high expression, and the blue curve represents low expression. (c) The survival analysis results of LRRC48 in different subgroups: the abscissa axis represents the survival time, and the ordinate axis represents the survival probability. The red curve represents high expression, and the blue curve represents low expression. (d) The survival analysis results of cg25726128 in different subgroups: the abscissa axis represents the survival time, and the ordinate axis represents the survival probability. The red curve represents high expression, and the blue curve represents low expression. (e) Immunotherapy efficacy score calculated by TMEscore group: the abscissa axis represents the TMEscore subgroup, and the ordinate axis represents TIDE. ∗∗ represents that the analysis result is statistically significant. (f) Using ROC analysis to evaluate the predictive ability of TMB, TMEgroup, and TMB+TME group on the effect of immunotherapy. (g) The relationship between MSI and TMEscore: the abscissa axis represents different MSI subgroups, and the ordinate axis represents TMEscore. ∗∗ represents that the analysis result is statistically significant. (h) Comprehensive genome landscape of BRCA (47 survival-related genes, P < 0.01).
|
PMC9071896
|
BMRI2022-5673810.006.jpg
|
0.435716 |
5a232d636ab149738c37e168fb740006
|
1H NMR spectra of the monomer.
|
PMC9071944
|
c9ra04828f-f1.jpg
|
0.424406 |
cacfa72d94d4496388863b788b8c8279
|
Results from nanoindentation tests for the cured resin.
|
PMC9071944
|
c9ra04828f-f10.jpg
|
0.527658 |
e84c6d2f30f14a2fabd5abf62eae3336
|
DMA curves of the cured resin.
|
PMC9071944
|
c9ra04828f-f11.jpg
|
0.514916 |
feebf9bcfcf641afbba37c0d7f1a1858
|
FT-IR spectra of the monomer and resin.
|
PMC9071944
|
c9ra04828f-f2.jpg
|
0.522294 |
33bd4cabe51a416b971b872b861a875e
|
DSC curve of Monomer 3 at a heating rate of 10 °C min−1 in N2.
|
PMC9071944
|
c9ra04828f-f3.jpg
|
0.52205 |
5f211f51f19f4808910ea1257465a1df
|
TG curve of the cured resin in N2 with a heating rate of 10 °C min−1.
|
PMC9071944
|
c9ra04828f-f4.jpg
|
0.517435 |
68ff41b5cfce4019883a85573d73aae0
|
DTG curve of the cured resin in N2 with a heating rate of 10 °C min−1.
|
PMC9071944
|
c9ra04828f-f5.jpg
|
0.447709 |
34656cadaf45438e9bf2cb2df0514582
|
Contact angle of water on the cured resin.
|
PMC9071944
|
c9ra04828f-f6.jpg
|
0.470885 |
89bb001c057c4a5987bd38d57a378962
|
AFM images of the cured resin film: (a) planar graph and (b) stereogram.
|
PMC9071944
|
c9ra04828f-f7.jpg
|
0.507214 |
b37ea1564a034505943e7a7c2842a42a
|
Dielectric constant and dielectric loss of the cured resin.
|
PMC9071944
|
c9ra04828f-f8.jpg
|
0.45041 |
1d3c36c63c7847ccb13a80562311ed43
|
X-ray diffraction (XRD) patterns of the cured resin (powder).
|
PMC9071944
|
c9ra04828f-f9.jpg
|
0.415888 |
9be29a28cef6468c9ec6b2217afd5b47
|
Reagents and conditions: (i) NaNO2, concentrated HCl, 0 °C, 1 h; HNO3 : H2O (3 : 2), 17 °C, 75 min; (ii) EtOH, reflux, 24 h; (iii) pyridine, 50 °C, 30 min.
|
PMC9072087
|
c9ra05712a-f1.jpg
|
0.428227 |
afa554f4fc7249d2b6d146aba4575d51
|
Optical microscope and scanning electron micrographs of the hyphae of S. sclerotiorum grown on PDA medium with DMSO or compounds 2, 5c at 25 °C. Optical microscope: (A) untreated control, 0.5% DMSO, ×400; (B) compound 2 at 0.0016 mM (EC50) treatment, ×400; (C) compound 5c at 0.0030 mM (EC50) treatment, ×400; scanning electron microscopy: (D) untreated control, 0.5% DMSO, ×1500; (E) after 72 h compound 2 at 0.0016 mM (EC50) treatment, ×1500; (F) after 72 h compound 5c at 0.0030 mM (EC50) treatment, ×1500.
|
PMC9072087
|
c9ra05712a-f2.jpg
|
0.371974 |
cf1a1aa5ec81441cab1d4491eb08cab7
|
In vivo protective efficacy of compounds 2, 5c and azoxystrobin against S. sclerotiorum on rape leaves.
|
PMC9072087
|
c9ra05712a-f3.jpg
|
0.536259 |
34d674f0972a42cab215f027ca7845bf
|
DSC of perdecanoic acid at a heating rate of 10 °C min−1.
|
PMC9072111
|
c9ra06087a-f1.jpg
|
0.448481 |
c1849a434a0c40049efaab6c878cf5ff
|
The activity of peracids in the model oxidation of 2-adamantanone. Reaction conditions: 2-adamantanone (0.100 g, 0.67 mmol), peracid (1.34 mmol), toluene (2 mL), 25 °C, 1200 rpm.
|
PMC9072111
|
c9ra06087a-f2.jpg
|
0.523724 |
baf7c9d370074904bcb40d09734d2605
|
The influence of the solvent for the model oxidation of 2-adamantanone. Reaction conditions: 2-adamantanone (0.100 g, 0.67 mmol), perdecanoic acid (1.34 mmol), solvent (2 mL), 25 °C, 1200 rpm.
|
PMC9072111
|
c9ra06087a-f3.jpg
|
0.443058 |
fa075fe9534e4f3abf528bd3be6f5133
|
The influence of the temperature for the model oxidation of 2-adamantanone. Reaction conditions: 2-adamantanone (0.100 g, 0.67 mmol), perdecanoic acid (1.34 mmol), toluene (2 mL), 1200 rpm.
|
PMC9072111
|
c9ra06087a-f4.jpg
|
0.476406 |
f9228813f2784f4f80d4f438bc8cd89c
|
The influence of molar ratio ketone : oxidant for the model oxidation of 2-adamantanone. Reaction conditions: 2-adamantanone (0.100 g, 0.67 mmol), perdecanoic acid, toluene (2 mL), 35 °C, 1200 rpm.
|
PMC9072111
|
c9ra06087a-f5.jpg
|
0.468649 |
a3e5042b9d8e40da86d95a1661248c76
|
(a) XRD pattern, TEM image (inset), (b) XPS spectrum and (c) proton longitudinal r1 and transverse r2 relaxivities (measured at 1.41 T and 37 °C) of pristine Fe3O4/γ-Fe2O3 nanoparticles.
|
PMC9072193
|
c9ra07227f-f1.jpg
|
0.486335 |
5cb2f938e59e42ac97ead1144f8f1451
|
Proton longitudinal r1 and transverse r2 relaxivities (measured at 1.41 T and 37 °C) of (a) Ca2+, (b) Fe3+, (c) Na+, (d) Mg2+, (e) Zn2+, (f) Ni2+, (g) Co2+, and (h) Cd2+ adsorbed Fe3O4/γ-Fe2O3 nanoparticles.
|
PMC9072193
|
c9ra07227f-f2.jpg
|
0.454902 |
15cc2705f2eb4d70b8d41e42ac8bbdaa
|
Dependence of (a) r1 and (b) r2 values on Xr/Z with regard to various cation-adsorbed Fe3O4/γ-Fe2O3 nanoparticles. The experimental data are plotted as solid squares, and the solid lines are fitting results according to eqn (1).
|
PMC9072193
|
c9ra07227f-f3.jpg
|
0.388584 |
88c5312fb369479fb1a7e6a7c902d54e
|
Dependence of (Xr/Z − xc)/w values for various cation-adsorbed Fe3O4/γ-Fe2O3 nanoparticles.
|
PMC9072193
|
c9ra07227f-f4.jpg
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.