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
|
---|---|---|---|---|
0.419077 |
bae422556be0465d84ab494b842ebe43
|
The influence of IOT grinding time on the (a) cumulative intrusion volume, and (b) pore volume distributions of the 28-d concrete.
|
PMC9181601
|
materials-15-03866-g009.jpg
|
0.418699 |
d27f6b39c2ed4c249bca005580aabe9d
|
The influence of SCMs content on the (a) cumulative intrusion volume, and (b) pore volume distributions of the 28-d concrete.
|
PMC9181601
|
materials-15-03866-g010.jpg
|
0.395195 |
4a2d389bb65e45ef8430299f3c83943b
|
Identified BSE images of samples at 28-d of (a) G0, (b) G1.5, (c) G2, (d) C0, (e) C20, and (f) C30.
|
PMC9181601
|
materials-15-03866-g011.jpg
|
0.509965 |
07ebf277cce24c5a855722399d604152
|
Effect of IOT grinding time on (a) Porosity, and (b) Anhydrous profiles of 28-d concrete.
|
PMC9181601
|
materials-15-03866-g012.jpg
|
0.41962 |
050c2c366c8747769acb9d10f693d2f8
|
Effect of SCMs content on (a) Porosity, and (b) Anhydrous profiles of 28-d concrete.
|
PMC9181601
|
materials-15-03866-g013.jpg
|
0.413329 |
5d9acda06441498782ce87926f536b1f
|
Several triazole drugs for IFIs in clinic and the structures of our lead compounds.
|
PMC9182106
|
molecules-27-03370-g001.jpg
|
0.424235 |
ab681cafd71143c1b598eba1ef734a55
|
Design strategy of target compounds.
|
PMC9182106
|
molecules-27-03370-g002.jpg
|
0.488303 |
e4adce84ec044febb27f7eee6ff5f892
|
The evaluation chart of Alog P and PSA of target compounds. The 95% and 99% confidence limit ellipses of blood–brain barrier (BBB) and human intestinal absorption (Absorption) models.
|
PMC9182106
|
molecules-27-03370-g003.jpg
|
0.378758 |
4bd75514131c4009b38ceb7ca6f1e465
|
Anti-hyphal effects of different concentrations of compounds 5k and 6c with FCZ (fluconazole) as a positive agent. Exponentially growing C. albicans SC5314 cells were transferred to hypha-inducing Spider liquid medium. The cellular morphology was photographed after incubation at 37 °C for 3 h. Scale bar = 20 μm.
|
PMC9182106
|
molecules-27-03370-g004.jpg
|
0.502806 |
1aabd9763b6643dfb11faa5f8a65b7bd
|
Effects of the antifungal treatments on the tissue burden. Each mouse was intravenously infected with C. albicans SC5314 5 × 106. Saline, FCZ (fluconazole, 0.3 mg/kg), compound 6c (0.3 mg/kg and 1.0 mg/kg) were administered orally once daily for 3 days after infection. Kidneys of mice were harvested on day 4 and colony-forming units were measured. * p < 0.05, ** p < 0.01, *** p < 0.001, determined by ANOVA.
|
PMC9182106
|
molecules-27-03370-g005.jpg
|
0.459056 |
b339f72fced14a8d8fae5e1992231360
|
In vivo efficacy of compound 6c (0.5, 1.0 and 2.0 mg/kg) and FCZ (fluconazole, 0.5 mg/kg) in a systemic infection of an ICR mouse model (n = 10) with C. albicans SC5314. ** p < 0.01, *** p < 0.001, determined by ANOVA.
|
PMC9182106
|
molecules-27-03370-g006.jpg
|
0.422658 |
4a52883878cc4793bd7ce72ae9c2a059
|
The binding mode of compound 6c in the active site of CYP51. Compound 6c, heme group, residues are shown in brown, yellow or green sticks, respectively. Hydrogen-bonding interaction was shown in red and π–π stacking interaction was shown in forest green. Image depicting the proposed binding mode was generated using PyMOL.
|
PMC9182106
|
molecules-27-03370-g007.jpg
|
0.457234 |
a185e87fe7164ebd9e37d007c89be901
|
Synthesis of the target compounds. (i) 4-Iodo-1H-pyrazole, K2CO3, DMF, 80 °C, 4 h; (ii) Methyl 4-ethynylbenzoate, PdCl2(PPh3)2, CuI, DIEA, NMP, 60 °C, 6 h; (iii) LiOH, THF/H2O (V/V= 1:1), 50 °C, 6 h; (iv) Alkyl amines or aromatic amines, DIEA, PyBOP, 50 °C, 4~8 h; (v) substituted (4-ethynylphenoxy)methyls, PdCl2(PPh3)2, CuI, DIEA, NMP, 60 °C, 6 h.
|
PMC9182106
|
molecules-27-03370-sch001.jpg
|
0.409737 |
903907db9fc04fe7ac72da17194f781a
|
Possible applications of ceramic nanofibers.
|
PMC9182284
|
materials-15-03909-g001.jpg
|
0.509011 |
90729976e44b4d308e62fb0c5619bd66
|
Main studies addressing ceramic nanofibers applied in wound-healing applications over the last six years [20,66,67,68,69,70,71].
|
PMC9182284
|
materials-15-03909-g002.jpg
|
0.473939 |
b4bfe8fb08d5432286c988b1358b6bb2
|
Evolution of the implant area in Wistar rats: (A) control rat prior to surgical intervention; (B) two weeks after material implantation, being observed inflammation in the incisions on the subcutaneous tissue; (C) four weeks of the surgical intervention, with a significant decrease in inflammation; and (D) six weeks after the intervention, the rat showed a very noticeable surgical decrease in the incisions’ inflammation, with a considerable growth in the rat’s hair and its scars (reprinted from Garibay-Alvarado et al. [74], copyright (2021), with permission from PloS ONE).
|
PMC9182284
|
materials-15-03909-g003.jpg
|
0.434144 |
d025275842774794a5ea9eaf15316745
|
Schematic representation of the influence of ceramic nanofibers on the wound healing process.
|
PMC9182284
|
materials-15-03909-g004.jpg
|
0.403193 |
b27c9265192440eb8ca4c5dd44a405b8
|
Optical images of an in vitro wound-healing assay undertaken on a human skin fibroblast cell line: (A) SV 40-transformed GM 00637; (B) scratch created by micropipette on the confluent cell culture plate; (C–E) control sample at 4, 6, and 24 h, respectively; (F–H) ABGnf (without boron) at 4, 6, and 24 h; and (I–K) ABGnf (with boron) at 4, 6, and 24 h, respectively. Reprinted from [88], copyright (2020), with permission from the International Journal of Applied Glass Science.
|
PMC9182284
|
materials-15-03909-g005.jpg
|
0.484221 |
06009b2ab7ac4b82908b0f6367aa4ecf
|
Main studies addressing ceramic nanofibers applied in bone regeneration applications over the last six years [23,55,56,74,114,115,116,117,118,119,120,121].
|
PMC9182284
|
materials-15-03909-g006.jpg
|
0.374951 |
96a088b90db949e1a0b5034b0c6f87ef
|
Live/dead immunofluorescence staining results of osteoblast cells cultured for one (A,D), three (B,E), and five (C,F) days on binary glass nanofibrous scaffold (A–C), as well as a blank control (D–F). Reprinted from Luo et al. [55], copyright (2017), with permission from Materials Science and Engineering: C.
|
PMC9182284
|
materials-15-03909-g007.jpg
|
0.477956 |
29c00a4ad7de412396fc7bfd356f8823
|
SEM images and the corresponding high magnification images of NBG scaffolds (A), after immersion in SBF for (B,C) one, (D,E) three, and (F,G) seven days (insets show local enlarged areas). Reprinted from [156], copyright (2018), with permission from RSC Advances.
|
PMC9182284
|
materials-15-03909-g008.jpg
|
0.443594 |
3eaf2a907c3b4c5ca5d195b186835c56
|
(A) MTT activity of MC3T3-E1 cells cultured on studied scaffolds for 7, 14, and 21 days. (B) Total protein content (μg/mL), up to 7, 14, and 21 days of cell culture of the osteoblastic lineage MC3T3 cultured on the studied scaffolds. (C) Alkaline phosphatase activity (U/L) of MC3T3-E1 cells on the studied scaffolds. (* statistical significant differences with p-value < 0.05; ** statistical significant differences with p-value < 0.01; *** statistical significant differences with p-value < 0.001). Reprinted from [23], copyright (2021), with permission from Ceramics International.
|
PMC9182284
|
materials-15-03909-g009.jpg
|
0.419088 |
2f0b5aaedce647c281e9aa3d32456652
|
(A) Weight loss, cumulative release of (B) calcium ions and (C) silicate ions from sintered nanofibers of CS-800, CS-1000, and CS-1200; (D1–D3) shows SEM images and (E1,E3) TEM images of the corresponding calcium silicate nanofibers, after degradation in deionized water at 37 °C for 21 days; (D1,E1) corresponding to CS-800 composition; (D2,E2) corresponding to CS-1000 and (D3,E3) to CS-1200. Reprinted from [170], copyright (2019), with permission from Ceramics International.
|
PMC9182284
|
materials-15-03909-g010.jpg
|
0.442086 |
58319f12fbee43aab16cb36d6f7f281c
|
Schematic representation of the influence of ceramic nanofibers on the wound-healing process.
|
PMC9182284
|
materials-15-03909-g011.jpg
|
0.455453 |
af04f6f5e6ce4bdea4bf3e70e26f941f
|
Evaluation of osteogenic differentiation of BMSCs cultured on doped and non-ion doped nanofibers compositions by an analysis of osteogenesis-related markers including: (A) quantitative analysis on ALP activity; (B) quantitative analysis on COL-I synthesis; (C) ALP staining and Alizarin red staining for calcium modules. * p < 0.05, significant; ** p < 0.01 and *** p < 0.001, highly significant (n = 4). Reprinted from [197], copyright (2021), with permission from Ceramics International.
|
PMC9182284
|
materials-15-03909-g012.jpg
|
0.453541 |
eb46188538b1451e962202028a0d9a02
|
The effect of CPF intervention on intestinal microbiota in mice with circadian rhythm disturbance: (A) Venn diagrams of the operational taxonomic unit (OTU) in the three groups. (B) Circos plot of the top5 microbial taxa at the phylum level. Microbial distributions of different groups at the phylum (C) and genus levels (D). Difference analysis of gut microbiota composition at the genus level between the CD and CT groups (E) and between the CD and CPF groups (F). (G) Dominant microbial taxa of the CPF group. (H) The alternation of the microbial taxa in the CD group. (I) The alternation of the microbial taxa in the CPF group.
|
PMC9182649
|
nutrients-14-02308-g001.jpg
|
0.453025 |
5e780a54c4e54f618542131f87050e79
|
The effect of CPFs on gut metabolites in mice with circadian rhythm disturbance. (A) The classification of differential metabolites in the CD group. (B) Differential metabolites between the CT and CD groups. (C) Differential metabolites between the CPF and CD groups. (D) KEGG enrichment pathways of significantly regulated differential metabolites between the CD and CT groups. (E) KEGG enrichment pathways of significantly regulated differential metabolites under CPF supplementation.
|
PMC9182649
|
nutrients-14-02308-g002.jpg
|
0.406579 |
f72eea4e41514fdc85bb4e695e40a8cf
|
Spearman analysis of the correlation between intestinal microbial composition and intestinal differential metabolites: (A) Spearman correlation analysis between genus and metabolite concentrations affected by CPF supplementation. (B) Network diagram of intestinal differential metabolites and microbial taxa. * p < 0.05.
|
PMC9182649
|
nutrients-14-02308-g003.jpg
|
0.448956 |
19bc1439794b46078ba801edc565cdcf
|
Single-cell RNA-seq identified circadian-rhythm-associated brain cell populations in the hypothalamus of mice: (A) t-SNE map of single cells of the hypothalamus in the CD, CT, and CPF groups. (B) t-SNE images of total hypothalamic cells in different clusters. (C) t-SNE identification map of hypothalamic single cells in the CT, CD, and CPF groups of samples. (D) t-SNE identification map of hypothalamic single cells in the CD group. (E) t-SNE identification map of hypothalamic single cells in the CPF group. (F) The percentage of cell types identified per cell cluster.
|
PMC9182649
|
nutrients-14-02308-g004.jpg
|
0.391722 |
7e5c531f5b444feea9f5526e6287550a
|
Effect of CPFs on the expression of circadian clock genes in the hypothalamus: t-SNE map of the expression of rhythm genes in the hypothalamus in the CPF group (A) and CD group (B). (C) Dot plots of circadian rhythm gene expression levels in different clusters of the CPF group. (D) Violin plot of the expression of circadian rhythm genes in the main cell types of the CPF group.
|
PMC9182649
|
nutrients-14-02308-g005.jpg
|
0.420232 |
5177376af3b0474ea0441078139120ac
|
Effects of CPF intervention on gene enrichment pathways in the hypothalamus: (A) GO classification analysis of DEGs in the hypothalamus between the CD and CPF groups. (B) The regulation of CPF intervention on DEGs involved in myelination of the CNS. (C) KEGG classification analysis of DEGs in the hypothalamus between the CD and CPF groups. (D) The regulation of CPF intervention on DEGs involved in insulin secretion. (E) The regulation of CPF intervention on DEGs involved in Parkinson disease. (F) The regulation of CPF intervention on DEGs involved in Alzheimer disease.
|
PMC9182649
|
nutrients-14-02308-g006.jpg
|
0.441542 |
b46e21492ee64382a457b8404e2729c1
|
Combined effect of field and year (F x Y) on bunch weight (a) and bunch number (b) of V. vinifera subsp. vinifera ‘Falanghina’ vines at the four study sites: SL-Santa Lucia, CA-Calvese, GR-Grottole, AC-Acquafredde. Mean values and standard errors are shown. Different letters indicate significant differences according to Duncan’s multiple range test (p ≤ 0.05).
|
PMC9182941
|
plants-11-01507-g001.jpg
|
0.362865 |
c2c5c9d4ecbd450e96a8ef4a1d68329d
|
Combined effect of field and year (F x Y) on net photosynthetic rate (Pn) of V. vinifera subsp. vinifera ‘Falanghina’ vines at the four study sites: SL-Santa Lucia, CA-Calvese, GR-Grottole, AC-Acquafredde. Mean values and standard errors are shown. Different letters indicate significant differences according to Duncan’s multiple range test (p ≤ 0.05).
|
PMC9182941
|
plants-11-01507-g002.jpg
|
0.471389 |
598bf982b059487e8ff262cd58644552
|
Epi-fluorescence microscopy views of abaxial leaf epidermis of V. vinifera ‘Falanghina’ vines at the four study sites: SL-Santa Lucia (a), CA-Calvese (b), GR-Grottole (c), AC-Acquafredde (d). Images are all at the same magnification. Bar = 50 µm.
|
PMC9182941
|
plants-11-01507-g003.jpg
|
0.430778 |
39d7789843864b6f88ba30dad2d66ea9
|
Combined effect of field and year (F x Y) on stomatal frequency of V. vinifera subsp. vinifera ‘Falanghina’ vines at the four study sites: SL-Santa Lucia, CA-Calvese, GR-Grottole, AC-Acquafredde. Mean values and standard errors are shown. Different letters indicate significant differences according to Duncan’s multiple range test (p ≤ 0.05).
|
PMC9182941
|
plants-11-01507-g004.jpg
|
0.442126 |
5a81056833f143fc9b2c80e4b02cad92
|
Combined effect of field and year (F x Y) on Total VAA (a) and FVEA (b) of V. vinifera subsp. vinifera ‘Falanghina’ vines at the four study sites: SL-Santa Lucia, CA-Calvese, GR-Grottole, AC-Acquafredde. Mean values and standard errors are shown. Different letters indicate significant differences according to Duncan’s multiple range test (p ≤ 0.05).
|
PMC9182941
|
plants-11-01507-g005.jpg
|
0.570067 |
caa199102c5048f9acfb320ea2e19745
|
The four experimental sites Santa Lucia (a), Calvese (b), Grottole (c), Acquefredde (d) vineyards.
|
PMC9182941
|
plants-11-01507-g006.jpg
|
0.522643 |
e9bef397fc7d41b6b50a87fa98786e13
|
Light microscopy views of V. vinifera ‘Falanghina’ leaf lamina sample with arrows pointing to the FVEA (2, second-order vein). Bar = 300 µm.
|
PMC9182941
|
plants-11-01507-g007.jpg
|
0.45277 |
4d1aad737b4d476480f96f491b8fff18
|
Circular symbiosis system in ecological production park.
|
PMC9184201
|
CIN2022-8410996.001.jpg
|
0.53863 |
94d628e048764ffaad2263aaa671ba7b
|
Ensemble learning optimization model based on parallel evolution method.
|
PMC9184201
|
CIN2022-8410996.002.jpg
|
0.443531 |
8fd99c2d2e1e4096837d0bbb555d8286
|
Structure of AG and AH production park.
|
PMC9184201
|
CIN2022-8410996.003.jpg
|
0.42711 |
a9e65fcbc002472d946f4da90f08265d
|
Experimental design and verification environment.
|
PMC9184201
|
CIN2022-8410996.004.jpg
|
0.434582 |
55dbabfe7d32481395a01bb76777e8e1
|
Water storage of different crop species in AG and AH production park before and after model optimization (1, rice; 2, wheat; 3, corn; 4, potato; 5, tomato; 6, cucumber).
|
PMC9184201
|
CIN2022-8410996.005.jpg
|
0.516385 |
2a86a9c36a8e4c5d839618be0041d79a
|
Water storage of different crop species in the experimental field before and after model optimization (1, rice; 2, wheat; 3, corn; 4, potato; 5, tomato; 6, cucumber).
|
PMC9184201
|
CIN2022-8410996.006.jpg
|
0.392204 |
23be231e4ede4abcb4a4821ed15ae7f5
|
Crop yield of different crop species before and after model optimization in the AG and AH production park (1, rice; 2, wheat; 3, corn; 4, potato; 5, tomato; 6, cucumber).
|
PMC9184201
|
CIN2022-8410996.007.jpg
|
0.393478 |
e5fb2277a194437c8847852a27dc9dd5
|
Crop yield of different crop species in the experimental field before and after model optimization (1, rice; 2, wheat; 3, corn; 4, potato; 5, tomato; 6, cucumber).
|
PMC9184201
|
CIN2022-8410996.008.jpg
|
0.451387 |
6bbf91bd2f614bc596732e6fcf2215fa
|
Crop survival rate of different crop species before and after optimization in the AG and AH park (1, rice; 2, wheat; 3, corn; 4, potato; 5, tomato; 6, cucumber).
|
PMC9184201
|
CIN2022-8410996.009.jpg
|
0.410761 |
011cf648042e4e61af00f0bc7dfdc22f
|
Crop survival rate of different crop species in the experimental field before and after optimization (1, rice; 2, wheat; 3, corn; 4, potato; 5, tomato; 6, cucumber).
|
PMC9184201
|
CIN2022-8410996.010.jpg
|
0.413224 |
4d83082fde2e406293561cdaca4bbd8d
|
Principle of compact light field photography.a A conventional light field camera captures the scene from different views with a lens array and records all sub-aperture images. In contrast, CLIP records (operator \documentclass[12pt]{minimal}
\usepackage{amsmath}
\usepackage{wasysym}
\usepackage{amsfonts}
\usepackage{amssymb}
\usepackage{amsbsy}
\usepackage{mathrsfs}
\usepackage{upgreek}
\setlength{\oddsidemargin}{-69pt}
\begin{document}$${{{{{{\bf{A}}}}}}}_{{k}}$$\end{document}Ak) only a few nonlocal measurements (\documentclass[12pt]{minimal}
\usepackage{amsmath}
\usepackage{wasysym}
\usepackage{amsfonts}
\usepackage{amssymb}
\usepackage{amsbsy}
\usepackage{mathrsfs}
\usepackage{upgreek}
\setlength{\oddsidemargin}{-69pt}
\begin{document}$${{{{{{\bf{f}}}}}}}_{{k}}$$\end{document}fk to fn) from each sub-aperture image and exploits the depth-dependent disparity (modeled by Bk) to relate the sub-aperture images for gathering enough information to reconstruct the scene computationally. Refocusing is achieved by varying the depth-dependent disparity model Bk. b Seeing through severe occlusions by CLIP as a camera array, with each camera only recording partial nonlocal information of the scene. A obscured object (from the camera with black rays) remains partially visible to some other views (with green rays), whose nonlocal and complementary information enables compressive retrieval of the object. c Illustration of instantaneous compressibility of the time-of-flight measurements for a 3D scene in a flash LiDAR setup, where a transient illumination and measurement slice the crowded 3D scene along the depth (time) direction into a sequence of simpler instantaneous 2D images. d–f CLIP embodiments that directly perform nonlocal image acquisitions with a single-pixel, a linear array, and 2D area detectors, respectively. A single pixel utilizes a defocused spherical lens to integrate a coded image, with u and v behind the lnes being the angular dimension. A cylindrical lens yields along its invariant axis a radon transformation of the en-face image onto a 1D sensor. The complex-valued mask such as a random lens produces a random, wide-field PSF that varies with object depth to allow light field imaging. PSF point spread function, CLIP compact light field photography, LiDAR light detection and ranging, 1D, 2D, 3D one, two, and three-dimensional.
|
PMC9184585
|
41467_2022_31087_Fig1_HTML.jpg
|
0.423435 |
b738b815c1ee44c78f1fd05701c604e5
|
Three-dimensional imaging (3D) through occlusions.a–c Reconstructed 3D images rendered in different perspective for three scenes: circular plate (a) and letter V (b) behind the letter N, and letter X (c) blocked by a rectangular plate. The severe occlusions are evident from the front view images, with the larger objects in the front completely blocked the object right behind them. In contrast, CLIP is able to unambiguously reconstruct the obstructed objects in 3D without any defocusing signals from the preceding occluder. d Three representative frames of imaging a 2 × 2 grid pattern moving across the CLIP camera FOV behind a rectangular occluder. Note that signals from the black occluders are enhanced relative to the objects for better visualization.
|
PMC9184585
|
41467_2022_31087_Fig2_HTML.jpg
|
0.426643 |
ea405944f2384409bde86df35da01331
|
Snapshot flash LiDAR imaging over an extended depth range.a Flash LiDAR imaging of a letter scene. From left to right are the reference photographs, a projected two-dimensional LiDAR images along the depth direction, and the 3D (three-dimensional) point-cloud representation of the scene. b flash LiDAR of the same 3D scene without extending the imaging depth of field, obtained by refocusing the camera onto a single focal plane. Note the defocus blur in the near and far objects. c Computational all-in-focus image. d and e Two representative frames for the dynamic imaging of a manually rotated letter V in a simple and cluttered scene, respectively.
|
PMC9184585
|
41467_2022_31087_Fig3_HTML.jpg
|
0.407349 |
2cd6417124d241cea6603f9889a38213
|
NLOS imaging by CLIP-ToF.a–c Imaging with planar, disconnected, and curved surfaces, respectively. From left to right are the flash LiDAR imaging of the relay surfaces, and two example hidden objects rendered as a projection image in the front view, and a 3D (three-dimensional) point cloud. Ground truth photographs of the object are shown in the inset of the front view image. d, e Reconstructed NLOS images for the disconnected and curved surfaces, respectively, with defocus errors on the relay wall, and those recovered with extended depth of field (highlighted by the green box). The quality of reconstruction degrades when the camera’s extended depth of field is disabled.
|
PMC9184585
|
41467_2022_31087_Fig4_HTML.jpg
|
0.41548 |
97f95e13afef41b584887b4c1308d6ec
|
Basic information of 3 groups of sensors: (a) Latitude and longitude range of each group of sensors; (b) Specific position of each group of sensors.
|
PMC9185465
|
sensors-22-04304-g001.jpg
|
0.453224 |
07ff568d66604b1ba3b6165fa1c857f4
|
Input data for three modes. (a) Input data of TR mode; (b) Input data of SR mode; (c) Input data of AR mode.
|
PMC9185465
|
sensors-22-04304-g002.jpg
|
0.367304 |
6878c63638fc49829028f0081f7960ca
|
Pearson correlation coefficient of input Matrix. (a) Correlation matrix of TR mode; (b) Correlation matrix of SR mode.
|
PMC9185465
|
sensors-22-04304-g003.jpg
|
0.428695 |
cd327ea789d54e5e9ec11ed78ccce8a9
|
Visualization results of AR data with t-SNE.
|
PMC9185465
|
sensors-22-04304-g004.jpg
|
0.498034 |
d6267441bac34ce6ab3448460e0fe039
|
Traditional FCM parameter selection result.
|
PMC9185465
|
sensors-22-04304-g005.jpg
|
0.406815 |
683c9f104c3844cba1388d39a1427d91
|
Example of TGO algorithm. (a) Data gridding; (b) Grid frequency distribution; (c) Cumulative contribution rate of each grid density; (d) Results of the first grid optimization; (e) Results of the second grid optimization.
|
PMC9185465
|
sensors-22-04304-g006.jpg
|
0.40949 |
805f37a2e8e548889688752d4c41feef
|
The selection result of thresholds.
|
PMC9185465
|
sensors-22-04304-g007.jpg
|
0.464969 |
4f5dcde0d7d44225af43f32487a17755
|
The XB index of each cluster center.
|
PMC9185465
|
sensors-22-04304-g008.jpg
|
0.41413 |
a411ba3fa3df4b33ad1e2143c0a43e62
|
Absolute error of different schemes at different loss ratios.
|
PMC9185465
|
sensors-22-04304-g009.jpg
|
0.399649 |
b4c2eeb6d7cf4314bdb668b80d5fa97c
|
The RMSE of different schemes on different groups of sensor data. (a) RMSE of the first group of sensors; (b) RMSE of the second group of sensors; (c) RMSE of the third group of sensors.
|
PMC9185465
|
sensors-22-04304-g010.jpg
|
0.421494 |
b71637442cc5473da1eaa55846ce0ae9
|
The RA of different schemes on different groups of sensor data. (a) RA of the No.1 sensors; (b) RA of the No.2 sensors; (c) RA of the No.3 sensors.
|
PMC9185465
|
sensors-22-04304-g011.jpg
|
0.471889 |
d05a628b885b4194a64cb69b6538da05
|
Comparison results of RMSE and RA.
|
PMC9185465
|
sensors-22-04304-g012.jpg
|
0.455241 |
43246553a4ba4e1e885b40ce6b5632a3
|
Effect of Assessment of Identity Development in Adolescence (AIDA) on BMI Increase over Time. Black lines indicate conditional regression lines for very low (M−2 SD), low (M−1 SD), medium (M), high (M +1 SD), and very high (M +2 SD) AIDA scores. BMI, Body mass index.
|
PMC9186337
|
fpsyt-13-887588-g0001.jpg
|
0.405436 |
254862362433483d99fd80855047f7a5
|
Effect of Assessment of Identity Development in Adolescence (AIDA) on BMI-SDS Increase over Time. Black lines indicate conditional regression lines for very low (M−2 SD), low (M−1 SD), medium (M), high (M +1 SD), and very high (M +2 SD) AIDA scores. BMI-SDS, Body mass index standard deviation score.
|
PMC9186337
|
fpsyt-13-887588-g0002.jpg
|
0.50654 |
1e8da102115c4fa6b6c941ba0acf6376
|
Flowchart of psychometric data analysis. ASQ:SE, Ages & Stages Questionnaires: Social-Emotional; KMO, Kaiser–Meyer–Olkin; NEST, Next Eigenvalue Sufficiency Test; PCA, principal component analysis.
|
PMC9187369
|
10-1055-s-0041-1741503-i2131588-1.jpg
|
0.442205 |
0508feb060da4a7295ec82ada07f5432
|
Screen plot with the exploratory factor analysis (EFA) and principal component analysis (PCA) results that were obtained with polychoric and Pearson correlations.
|
PMC9187369
|
10-1055-s-0041-1741503-i2131588-2.jpg
|
0.441568 |
000aec4169044a539e004286ba2d47a0
|
Graphical representation of exploratory factor analysis (EFA) and principal component analysis (PCA) models and network analysis results.
|
PMC9187369
|
10-1055-s-0041-1741503-i2131588-3.jpg
|
0.408445 |
342547f081184126b972271aaff02038
|
Nickel-catalyzed arylalkylation of alkynes.a The advantages of photocatalysis and electrochemistry. b Limitation of the electrochemical oxidation pathway and potential of the electrochemical reduction pathway in the cross-coupling field. c Initial results of nickel-catalyzed reductive cascade cross-couplings via the traditional reductant, electrochemical, and photocatalyzed pathways. d Stereocontrol of alkenes by switching between electrochemistry and photocatalysis. TDAE, tetrakis(dimethylamino)-ethylene; DMA, dimethylacetamide; PMDTA, N,N,N′,N′′,N′′-pentamethyldiethylentriamin; TMEDA, tetramethylethylenediamine; bpy, 2,2′-bipyridine.
|
PMC9187637
|
41467_2022_30985_Fig1_HTML.jpg
|
0.494542 |
b02b61ae55fe43a9b779e6c219c78212
|
Scope of the nickel catalyzed electrochemical arylalkylation of alkynes.Reactions were performed with 0.2 mmol of aryl bromides, 0.4 mmol of alkynes, and 0.6 mmol of alkyl bromide. Yields are of the isolated products.
|
PMC9187637
|
41467_2022_30985_Fig2_HTML.jpg
|
0.391662 |
2015304490b24203b1327f870a6ac00e
|
Scope of the photoredox/nickel dual catalyzed arylalkylation of alkynes.Reactions were performed with 0.4 mmol of aryl bromides, 0.2 mmol of alkynes, and 0.6 mmol of alkyl bromide. Yields are of isolated products. The values in parentheses represent the ratio of the two isomers (Psyn:Panti), which was determined by 1H NMR analysis.
|
PMC9187637
|
41467_2022_30985_Fig3_HTML.jpg
|
0.451572 |
f93df2e2395e44f498954490770d2dd9
|
Mechanistic study and scope of the photo-assisted electrochemically nickel catalyzed arylalkylation of alkynes.a Isomerization control reactions. b UV/Vis absorption spectroscopy and fluorescence emission spectroscopy of E and Z isomers. c Reactions were performed with 0.2 mmol of aryl bromides, 0.4 mmol of alkynes, and 0.6 mmol of alkyl bromide. Yields are of the isolated products. The values in parentheses represent the ratio of the two isomers (Psyn:Panti), which was determined by 1H NMR analysis. *Reaction performed in DMSO. #Reaction conditions: aryl bromides (0.2 mmol, 1 equiv.), alkynes (0.4 mmol), alkyl bromides (0.6 mmol), NiBr2·d(4-OMe)-bpy (10 mol%), TMEDA (3 equiv.), nBu4NBr (2 equiv.) in 4 mL DMA and electrolysis for 16 h at 4 mA using a graphite anode and nickel foam cathode. The reaction mixture was then directly irradiated with 2*390 nm purple LEDs for 24 h.
|
PMC9187637
|
41467_2022_30985_Fig4_HTML.jpg
|
0.438023 |
c3d25404f7514248bb55fb7906a99699
|
Applications.a Scale up reactions for both the E and Z isomers. b Complex molecule synthesis via three different approaches. c Further functionalization of the generated trisubstituted alkene.
|
PMC9187637
|
41467_2022_30985_Fig5_HTML.jpg
|
0.427407 |
e98d429e008f4142b28e9286506a5d71
|
Framework: construction of imaginations and the shaping of social practices and sociomaterial structures
|
PMC9188356
|
11625_2022_1161_Fig1_HTML.jpg
|
0.450486 |
c6d9dfc873d8403fa91745fd547d159f
|
Proximity of interviewees to trajectories (preservation,modernization, transformation) and scope of change
|
PMC9188356
|
11625_2022_1161_Fig2_HTML.jpg
|
0.430561 |
fc6432a983304cf0a74a82893c41f2df
|
Schematic illustration of the preparation of multi-layer mineralized GO-Col-HAp microgels with uniform HAp deposition and their applications in rat cranial defect and mandibular defect repair.
|
PMC9189211
|
gr1.jpg
|
0.477119 |
e044507b4730473a805c42aa77ad037f
|
Fabrication of multi-layer mineralized GO-Col-HAp microgels and their schematic illustrations. (A) Experimental steps for the synthesis of multi-layer mineralized GO-Col-HAp microgels using microarray chip and their chemical reaction illustration. (B) Schematic drawing, optical image, and micro ct evaluation of the microgels (Scale bars of low magnification: 500 μm, scale bars of high magnification and micro ct: 200 μm, red arrowheads refer to pore structure). (C) Schematic illustration, optical image, and micro ct analysis of the pearl (Scale bars of low magnification: 5 mm, scale bars of high magnification and micro ct: 1 mm). (D) The porosity of the microgels and peral. (E–F) Quantitative analysis of apatite volume to total volume (AV/TV) and apatite mineral density (AMD) of microgels and pearl (ns: no significance, ∗∗∗p < 0.001).
|
PMC9189211
|
gr2.jpg
|
0.455984 |
f7b49f0bda29428a9abf5887c74d7529
|
Morphology and physiochemical evaluation of the microgels. (A) SEM images of the surface, coronal-section, and cross-section morphologies of the microgels (Scale bars in low magnification images were 100 μm, and in high magnification images were 10 μm. The red regions represented for the inner layer, the brown regions represented for the middle layer, the blue regions represented for the outer layer, and the white arrowheads represented for HAp). (B) Schematic illustration and EDS mapping images of Ca and P elements and EDS point images of the MMGCH microgels. (C–E) FTIR spectra, XRD spectra, and TGA analysis of the microgels. (F) Pore size of the microgels. (G) Mass increase of the microgels after biomimetic mineralization in SBF. (H) Calcium quantitative analysis of the microgels. (I) Elastic modulus evaluation of the microgels (ns: no significance, ∗p < 0.05, ∗∗p < 0.01, ∗∗∗p < 0.001).
|
PMC9189211
|
gr3.jpg
|
0.400092 |
35e55b5a3632462eb438ae8855ab6f10
|
In vitro evaluation of migration and proliferation of r-BMSCs on/in the microgels. (A) Schematic illustration of cell migration and proliferation on/in the microgels. (B) Live/dead staining and SEM images of r-BMSCs on/in the microgels (Scale bars in confocal images: 100 μm, and scale bars in SEM images: 10 μm). (C) Quantitative analysis of living cells on/in the microgels at day 1, 4, and 7 post-incubation. (D) Cell viability of r-BMSCs on/in microgels at day 1, 4, and 7 post-incubation. (E) Quantitative analysis of fluorescence intensity in different layers measured at day 7. (∗p < 0.05, ∗∗∗p < 0.001, ns: no significance).
|
PMC9189211
|
gr4.jpg
|
0.442062 |
34505b9205834c478387df9a88eea70c
|
In vitro evaluation of r-BMSCs osteogenic differentiation on/in the microgels. (A) Schematic illustration of cell osteogenic differentiation on/in the microgels after 4 and 7 days of osteogenic incubation. (B) The protein expression of osteogenic markers of r-BMSCs. (C) Quantitative analysis of ALP, OCN, and COL-1 protein expression for 4 and 7 days on/in microgels, normalized to GAPDH. (D) qRT-PCR analysis of ALP, OCN, and COL-1 gene expression of r-BMSCs during 4 and 7 days of osteogenic incubation respectively. (E) OCN staining of microgels at day 4, 7 post-osteogenic incubation (red arrowheads: positive brown staining, scale bars: 100 μm, ns: no significance, ∗∗P < 0.01, ∗∗∗P < 0.001).
|
PMC9189211
|
gr5.jpg
|
0.375254 |
18df3714123343cc8a0998a07f867cae
|
In vivo evaluation of the microgels for critical-sized cranial and mandibular defect treatment in rat. (A–D) Representative 3D reconstruction and cross-section images of cranial and mandibular bone defects at week 4 and 12 post-implantation. (E, H) Optical images of the microgels in the defect regions during operation. (F–G) Quantitative evaluation of bone volume to tissue volume (BV/TV) and bone mineral density (BMD) in cranial bone defects. (I–J) Quantitative evaluation of BV/TV and BMD in mandibular bone defects. (ns: no significance, ∗P < 0.05, ∗∗∗P < 0.001).
|
PMC9189211
|
gr6.jpg
|
0.41228 |
f45a6a34769542559b2afa73a3c78ed8
|
Histological and immunohistochemical analysis of microgels after implantation for 4 and 12 weeks in cranial bone defects. (A) HE staining images of microgels complexes (Scale bars: 500 μm, white rectangle: the lateral margin and the center region of the defects). (B–E) HE staining (B), Masson's trichrome staining (C), OCN staining (D), and COL-1 staining (E) images of the microgels complexes engineered bone constructs (Scale bars: 100 μm, ∗: new bone, &: microgels).
|
PMC9189211
|
gr7.jpg
|
0.476818 |
4582320612864e8a9370fbc7c01cde77
|
Histological and immunohistochemical analysis of microgels after 4 and 12 weeks implantation in mandibular bone defects. (A) HE staining images of the defect region at week 4 and 12 pos-timplantation (Scale bars: 500 μm, white rectangle: the lateral margin and the center region of the defects). (B–E) HE staining, Masson's trichrome staining, OCN staining, and COL-1 staining images of lateral margin and center region of the defect region (Scale bars: 100 μm, ∗: new bone, &: microgels).
|
PMC9189211
|
gr8.jpg
|
0.492169 |
3c2a0a6ecb854a4cac742986e98ce05d
|
FAMYLY study design. Patients eligible for the study were treated with fish oil emulsion, daunorubicin and cytarabine at the indicated doses. If the baseline white blood count (WBC) was below 30 G/L, patients received 48 of fish oil before the start of chemotherapy. If the WBC was above or equal to 30 G/L, all treatments were started on the same day. An evaluation bone marrow (BM) evaluation was performed on D15. If the BM blasts were higher than 5%, a second induction was provided at the indicated doses.
|
PMC9192636
|
41598_2022_13626_Fig1_HTML.jpg
|
0.402498 |
7904b7903f924640b10bed285648c67c
|
PUFA plasma pharmacokinetics. Plasma samples were drawn at baseline (H0), 24 and 48 h after the start of FO infusion, after the end of infusion (D10), and at disease evaluation (D28–D35). The PUFA composition of plasma was evaluated by gas chromatography. The sum of docosahexaenoic acid (DHA) and eicosapentaenoic (EPA) acid concentration is indicated for each individual patient (colored lines and dots). Means are indicated by a dash for each timepoint.
|
PMC9192636
|
41598_2022_13626_Fig2_HTML.jpg
|
0.45016 |
6928aad02ee348ecb02e8e980f3f4a5f
|
Historical comparison of survival data from the FAMYLY and LAM2001 trials. Overall survival (OS) was plotted from the inclusion date to the date of death or last follow up. The vertical marks indicate censored observations.
|
PMC9192636
|
41598_2022_13626_Fig3_HTML.jpg
|
0.49276 |
f1f862ce551648cb842bfd585dcc0832
|
Effect of well-watered (WW) and terminal drought (TD) treatments on (A) total shoot biomass, (B) net photosynthesis rate, (C) photosynthetic efficiency of PSII and (D) instantaneous water use efficiency (WUEi) of two common bean (PS and BM) and one tepary bean genotypes (“Tep32”). Data are presented as the mean ± standard error of six biological replicates. Asterisks indicate significant differences between genotypes in the same water treatment by t-test. *p < 0.05, **p < 0.01, and ns not significant. Letters indicate statistical significance between genotypes determined with Tukey’s test. The statistical analyses between water treatments are not shown.
|
PMC9194640
|
fpls-13-894657-g001.jpg
|
0.455586 |
4533e4cbd5754b14ad74df69d330b91e
|
Stomatal density and epidermal cell density of two common bean genotypes (PS and BM) and one tepary bean (“Tep32”) under well-watered (WW) and terminal drought (TD) treatments in three different strata: basal (A), middle (B), and apical (C). Data are presented as the mean ± standard error. Inset: micrographs of representative epidermal impressions with stomata in blue for clarity. Scale bars = 50 μm.
|
PMC9194640
|
fpls-13-894657-g002.jpg
|
0.478648 |
9198e64a5d784bb9a1f15593fb4634df
|
Stomatal index (SI) and leaf area (left) and stomata width and stomata length (right) of two common bean genotypes (PS and BM) and one tepary bean (“Tep32”) under well-watered (WW) and terminal drought (TD) conditions in three different strata: basal (A), the middle (B), and apical (C). Data are presented as the mean ± standard error.
|
PMC9194640
|
fpls-13-894657-g003.jpg
|
0.424582 |
56112710bbb0445a9957f96904f97cb7
|
Transpiration efficiency of two common bean (PS and BM) and one tepary bean (“Tep32”) genotypes, under well-watered (WW) and terminal drought (TD) treatments. Data are presented as the mean ± standard error of six biological replicates. “ns” corresponds to not significant (p < 0.05).
|
PMC9194640
|
fpls-13-894657-g004.jpg
|
0.404252 |
50e9b191f3df48eb84368eff6ec6d050
|
Normalized transpiration rate vs. fraction of transpirable soil water of two common bean genotypes (PS and BM) and one tepary bean (“Tep32”). The solid line in each graph is the regression fit using the inverse exponential model. The dashed lines are the results of the two-segment plateau regression. Letters in the value of FTSW threshold indicate statistical significance between genotypes determined with Tukey’s test.
|
PMC9194640
|
fpls-13-894657-g005.jpg
|
0.401177 |
d313581bdcb04106ab8644078aa223e9
|
Effect of uncontrolled autophagy on plasma membrane‐associated proteins
A–ESurface views (A–D) and z‐sections (E) of third instar larval epidermis expressing the indicated fluorescent markers and RNAi or overexpression constructs. (A) Bazooka/Par3, normally seen at the apical adherens junction and in a perinuclear position, is lost or reduced in large areas under uncontrolled autophagy as are lateral membrane markers, (B) Fasciclin‐III (FasIII), (C) neuroglian (Nrg) and (D) adherens junctions (DE‐cad), whereas large bundles decorated with MRLC appear. (E) The baso‐lateral transmembrane protein β‐integrin is not lost; in extreme conditions, it is seen in both the apical and basal membranes (overexpression of Atg16B). Arrows in the z‐sections point to high accumulation of integrin along the folded lateral membranes.FElectron micrograph of a section through the larval epidermis to show the highly folded lateral junction between two epidermal cells (magenta and cyan). Left, cartoon; middle, false colouring; right, original image. C: cuticle; M: mitochondrion; BL: basal lamina; AJ: adherens junction; SJ: septate junction; GJ: gap junction, N: nucleus, Mv: apical microvilli connecting the cuticle to the cell.
Data information: A–E, n = 15–40 larvae each genotype. Scale bars: A–E, 20 μm and F, 2 μm. Images from Movies EV10 and EV11.
|
PMC9194749
|
EMBJ-41-e109992-g001.jpg
|
0.388512 |
05de0beec3844536a506b8527a54ba4b
|
Colocalization of Atg8a and septate junction component FasIII during epidermal wound healing
A, BTime‐lapse series of single‐cell wound healing in a larva expressing mCherry‐Atg8a (magenta) (A58>mCherry‐Atg8a) and endogenously tagged FasIII (GFP gene trap; green), a transmembrane component of septate junctions. (B) Higher magnification of the post‐wounding time points in (A). Images from Movie EV18; see also Movies EV19 and EV20. Each frame is a merge of 68 planes spaced 0.28 μm apart. A, n = 11 larvaeCTransmission electron microscopy of larval epidermis overexpressing Atg16B shows autophagosomes in direct contact with plasma membrane (magenta arrows) and at the sites of cell–cell junctions (orange arrows). The lower panel shows a higher magnification of Atg16B
image in Fig 3D.
Data information: Scale bars: A, B 20 μm; C 500 nm.
|
PMC9194749
|
EMBJ-41-e109992-g002.jpg
|
0.382996 |
4746c73cdbbd43b0b68696c70459a1eb
|
Loss of lateral membrane integrity
A–DFluorescence loss in photobleaching (FLIP) to test free cytoplasmic GFP motility within the epidermis. A small area (magenta circle; 179 or 1,098 µm2) was laser‐illuminated for the indicated times (3 or 20 min) in control or Atg1‐expressing epidermis also expressing free GFP. (A) Snapshots before bleaching and at 3 points of recovery. (B) Kymographs along the broken line in (A) during recovery. (C, D) Quantification of fluorescence recovery after bleaching, shown separately for 3 and 20 min bleaching protocols.E, FElectron micrographs showing morphological defects or absence of lateral cell membranes in epithelia with upregulated autophagy. (E) Membrane domains with tight apposition between neighbouring cells are marked in magenta (adherens junctions), green (septate junction) and orange (gap junctions). (F) Nuclei (yellow) and lateral membranes in healthy epidermal cells cannot be shown in one image because they are too far apart, whereas when autophagy is upregulated, nuclei are often found close together and not separated by plasma membranes. C: cuticle (blue).
Data information: A–D n = 5–9 larvae for each FLIP protocol. Scale bars: A, 20 μm and E, F, 2 μm. Images from Movies [Link], [Link], [Link].
Source data are available online for this figure.
|
PMC9194749
|
EMBJ-41-e109992-g003.jpg
|
0.460009 |
a678a601953547eb91d9e18994d2ccad
|
Effect of uncontrolled autophagy on nuclear morphology
The nuclear lamina is visualized using a GFP‐tag inserted into the endogenous locus of the Kugelkern (Kuk) gene in animals expressing the indicated overexpression or RNAi constructs in the epidermis.A GFP‐tagged transgenic construct of Kugelkern is co‐expressed with the indicated overexpression or RNAi constructs. High levels of Kuk induce lobulation and other nuclear defects, which are ameliorated if TOR is downregulated or autophagy upregulated, but not if other branches of the TOR signalling pathways (S6K or rictor) are modified.
Data information: The lower rows show higher magnification of the nuclei marked above. n = 20–30 larvae each genotype. Scale bars: A, B, upper rows, 20 μm; lower rows, 5 μm.
|
PMC9194749
|
EMBJ-41-e109992-g004.jpg
|
0.388913 |
b00b416e2cb541ab9a361d96919f3651
|
Effect of uncontrolled autophagy on gut enterocytes and barrier function
Larval gut epithelia in which enterocytes express a marker for polarity and septate junctions, GFP‐Dlg using the NP1‐Gal4 driver together with or without Atg16B. Expression of Atg16B leads to disruption of lateral plasma membrane (magenta arrows). n = 32–44 larvae each genotype.Smurf gut barrier assay in anaesthetized, live larvae. Left, control larvae, right, larvae overexpressing Atg16B in the gut enterocytes under control of the NP1‐Gal4 driver. While the distribution of food in the gut of Atg16B‐expressing larvae was different from controls, there was no leakage of dye from the gut in any of the animals in three independent experiments (each with n = 30 larvae for each genotype).
Data information: Scale bars: A, 20 μm; B, 1,000 µm.
|
PMC9194749
|
EMBJ-41-e109992-g005.jpg
|
0.372882 |
ae709aba14de4491938ebacec8d2b6a8
|
Syncytium formation during wound healing
Schematic for timing of transgene expression and start of live imaging. Gene expression was induced at the end of the second larval instar, laser ablation and imaging started 30 h later.Wound healing in epithelia with clonally expressed cytoplasmic GFP (magenta) under the actin5c‐Gal4 driver in control larvae (act>GFP) or larvae expressing RNAi constructs specific for Atg1 or Atg5. Laser‐ablated cells are marked with white asterisks. To visualize cell borders DE‐cad‐RFP (green) was expressed in all tested genotypes. By the end of wound closure, GFP from the clonal cells has spread to all cells around the wound in the control, but not if autophagy is suppressed, regardless of the number of cells initially expressing GFP.Control larva in which a GFP‐expressing cell was wounded. No GFP is induced in or taken up by the surrounding cells.Control experiment in which the central cell was damaged but not killed (white marked area). No wound response occurs and no GFP leakage between neighbouring cells is seen.
Data information: (A–D) n = 6–9 larvae each genotype. The control pre‐wounding small clone in b (top left panel) is from Kakanj et al,
2016. B, C, D, z‐projections of time‐lapse series. B–D, Scale bars, 20μm. Pre W: pre‐wounding. Images from Movies [Link], [Link], [Link], [Link], [Link].
|
PMC9194749
|
EMBJ-41-e109992-g006.jpg
|
0.386753 |
406d6840ac554fddb0b4fb6c725ba9fd
|
Autophagy in unwounded epidermis
A–EControl of epidermal autophagy by TOR signalling. (A) Epidermis of third instar larvae expressing the autophagosome marker GFP‐Atg8a together with constructs for up‐ or downregulating the autophagy pathway in the epidermis. Healthy epidermis contains few autophagosomes, but artificially activating autophagy through overexpression of Atg1 or blocking TOR signalling leads to accumulation of autophagosomes. (B) Higher magnification of the areas marked by magenta boxes in (A). (C) Quantification of Atg8a puncta in an area of 10,000 µm2, n = 6–8 larvae each genotype. We assumed unequal sample size and unequal variances and calculations were performed. Values are presented as box plots. Box plot elements are: centre line, median; box limits, upper and lower quartiles; whiskers, 1.5× interquartile range; points, outliers. For statistical hypothesis testing, independent and non‐parametric (Kruskal–Wallis) t‐tests were performed for the mean number of spots in control and experimental conditions. P‐values are indicated as follows: *P < 0.04; **P < 0.003; ***P < 0.0002 and lack of an asterisk means non‐significant (P > 0.123). (D, E) Transmission electron micrographs of sections through the epidermis of larvae with elevated autophagy at two different magnifications. AP: autophagosome; Ax: cross‐section of an axon; BL: basal lamina; C: cuticle; HD: hemidesmosome; M: mitochondrion; N: nucleus.FDisruption of epidermal morphology after upregulation of autophagy. The membrane marker Src‐GFP is lost from large areas of the epidermis and nuclei have lost their regular spacing. Image in control and rictori
panels are from the time laps in Appendix Fig S5B and Movie EV12.GSchematic for temporally controlled transgene expression and imaging in (H). Gene expression is induced at the end of the second larval instar, live imaging started 6 h later and continued for an additional 6 h.HExample of membrane dynamics after time‐controlled Atg16B expression. Src‐GFP containing material appears to be taken out of and eventually detached from lateral cell membranes (arrows). Over the period of observation, abnormal accumulation of GFP is seen in the nuclei (cyan asterisks). t = 0 is 6 h after A58‐Gal4 activation. Image from Movie EV9.
Data information: A, B, F, H, z‐projections; n = 20–50 larvae each genotype. Scale bars: A, F, H, 20 μm, B, 10 μm, D, 1 μm and E, 500 nm.
Source data are available online for this figure.
|
PMC9194749
|
EMBJ-41-e109992-g007.jpg
|
0.411007 |
69e997fedc8142b19dd99c9e095555e0
|
Autophagy during epidermal wound healing
A–DAppearance of autophagic structures (marked with GFP‐Atg8a) during wound closure in the epidermis of control third instar larvae and after epidermal knockdown of the autophagy pathway components Atg1, Atg5, Atg6, Atg7 or Atg12. All constructs are expressed in the epidermis under the control of the A58‐Gal4 driver. (A) Time points from movies of wounded epidermis. The wounds have closed by 2 h in all cases. (B) Higher magnification of the areas marked by magenta boxes at t = 120 min. (C) Quantification of the appearance of GFP‐Atg8a puncta in the imaged area (10,000 µm2) 3 control larvae, shown for each individual larva. (D) Quantification of number of GFP‐Atg8a puncta in different genetic conditions measured in an area of 10,000 µm2 at the time of wound closure; n = 4–10 larvae each genotype, for the detail see Data analysis. We assumed unequal sample size and unequal variances and calculations were performed. Values are presented as box plots. Box plot elements are: centre line, median; box limits, upper and lower quartiles; whiskers, 1.5× interquartile range; points, outliers. For statistical hypothesis testing, independent and non‐parametric (Kruskal–Wallis) t‐tests were performed for the mean number of spots in control and experimental conditions. P‐values are indicated as follows: *P < 0.04; ***P < 0.0002 and lack of an asterisk means non‐significant (P > 0.123).E, FEffect of suppressing autophagy on wound healing and actin cable formation. Time‐lapse series of single‐cell wound healing in larvae expressing (E) Src‐GFP (green) and DsRed2‐Nuc (magenta) to mark cell membrane and nuclei and (F) endogenously GFP‐tagged E‐cadherin (DE‐cad‐GFP; green) and mCherry‐marked myosin regulatory light chain (Sqh‐mCherry; magenta) to visualize adherens junctions and actomyosin cables.
Data information: (A, B, E, F) Z–projections of time‐lapse series in early L3 larvae, n = 9–15 larvae each genotype. Scale bars, A, E, F, 20 μm and B, 10 μm. Pre W: pre‐wounding. Images from Movies [Link], [Link], [Link], [Link], [Link]. Genotypes of all images are listed in Table 2.
Source data are available online for this figure.
|
PMC9194749
|
EMBJ-41-e109992-g010.jpg
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.