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
0.53244
e9a8ec352ef64596b75bfc97846c7571
Total Smoke Production versus time for all the samples.
PMC10007242
polymers-15-01093-g017.jpg
0.459282
4507280eeb8941a29597d4f60153cc32
Flowchart of the preparation of phosphate containing glycerol.
PMC10007242
polymers-15-01093-sch001.jpg
0.467006
fb986fa18ccb42c8ba6f8a3d4ac32ebe
Flowchart of the preparation of phosphate containing glycerol.
PMC10007242
polymers-15-01093-sch002.jpg
0.498427
5bf2abe22bc948d4952b96543e14ee40
Flowchart of the preparation of FR-containing particleboards.
PMC10007242
polymers-15-01093-sch003.jpg
0.511127
d13f09e029d742a3aee159519f1b514f
The overall framework of online hashing preserving local features and global-balanced similarity (LSOH). (a) Manifold learning preserves the structural features of newly arrived data, and obtains the hash codes of newly arrived data through Laplacian Eigenmaps (LE). (b) Learn binary codes by a balanced similarity matrix built from newly arrived data and existing data to keep all the hash codes consistent. (c) Our proposed algorithm can learn hash codes preserving dual-semantic information, and obtain satisfactory retrieval results.
PMC10007520
sensors-23-02576-g001.jpg
0.467784
9ec544483a7f4844858d7cefd364c2ea
Example images of CIFAR-10 dataset.
PMC10007520
sensors-23-02576-g002.jpg
0.444817
a8c4f45ce4bb4f8e97df834ab7072575
Example images of MNIST dataset.
PMC10007520
sensors-23-02576-g003.jpg
0.417887
1dccfc267c4e4ad99ab85c5e9ec8b1a8
Example images of Places205 dataset.
PMC10007520
sensors-23-02576-g004.jpg
0.535143
c43092014b4044a9948b4c40b2c9987b
Comparisons of Precision@K curves on CIFAR-10.
PMC10007520
sensors-23-02576-g005.jpg
0.512728
bc59718ebfab40788a44626f795bb49b
Comparisons of Precision@K curves on MNIST.
PMC10007520
sensors-23-02576-g006.jpg
0.418587
e583809373f14b9f8cb7e594e9010400
PR curve under 32-bit hash codes. (a) PR curve on CIFAR-10 (b) PR curve on MNIST.
PMC10007520
sensors-23-02576-g007.jpg
0.488653
b524540717ee453abe9f414537ee0bc8
Comparisons of mAP(@1000) performances concerning varying values of αt, βt, γt when the hashing bit is 32.
PMC10007520
sensors-23-02576-g008.jpg
0.553429
b58aabc892ba4e28b22df6d6782840b5
X-ray diffractogram patterns with their corresponding Rietveld refinements of ZnO-Pure, ZnO-La, ZnO-Ce and ZnO-LaCe NPs performed on the MAUD software. Note: *, Secondary peaks.
PMC10009547
gr1.jpg
0.406014
66e0171be10a468d9562184be42a717a
TEM images of (a) ZnO-Pure (Al Bitar et al., 2022), (b) ZnO-La, (c) ZnO-Ce and (d) ZnO-LaCe NPs. The inset of (a) shows the average particle size distribution histogram for ZnO-Pure sample (Al Bitar et al., 2022). The specified scales (a) 50 nm, (b)100 nm, (c) 50 nm, and (d) 100 nm.
PMC10009547
gr2.jpg
0.487475
51c5a434b66f459fbfda488149929367
Photoluminescence emission spectra of ZnO-Pure (Al Bitar et al., 2022), ZnO-La, ZnO-Ce and ZnO-LaCe NPs.
PMC10009547
gr3.jpg
0.399608
17c7fae431c647818adaccbb090b1efe
Inhibitory effect of ZnO-Pure (Al Bitar et al., 2022), ZnO-La, ZnO-Ce and ZnO-LaCe, Ciprofloxacin, Doxycycline and Amoxicillin toward six bacteria, (a) Escherichia coli, (b) Klebsiella pneumonia, (c) Citrobacter braakii, (d) Staphylococcus haemolyticus, (e) Staphylococcus aureus, and (f) Streptococcus intermedius. Results are represented as mean ± SEM of at least three experiments. Abbreviation: SEM, standard error of the mean.
PMC10009547
gr4.jpg
0.467349
752393c0ac904ae6bc83d9962756737e
Effect of ZnO-Pure, ZnO-La, ZnO-Ce and ZnO-LaCe NPs on the cell viability of Caco-2 and HCT-116 cells. MTT assay was performed to detect the living cells after 24 or 48 h, respectively. The IC50 values were calculated after treating the cells with varying concentrations of the prepared NPs. The obtained data are represented as mean ± SEM of six independent replicates. Abbreviation: SEM, standard error of the mean.
PMC10009547
gr5.jpg
0.407013
1ce7d7a385624cf1865f954556244bee
Morphological changes induced by ZnO-Pure, ZnO-La, ZnO-Ce and ZnO-LaCe NPs with varying concentrations on Caco-2 and HCT-116 cells after 24–48 h using the Inverted Microscope with × 200 magnification with scale bar value 100 µm.
PMC10009547
gr6.jpg
0.446243
9bd67850d062439fb107c9187a9bbead
The PADMR spectrometer. (a) Lab-frame view of the PADMR spectrometer. The sample is mounted to a PCB loop antenna, inside an optical cryostat, between the pole tips of an electromagnet. The probe beam passes through the center of the RF antenna, and its intensity is measured by a photodetector. A laser is overlapped with the probe. For PA experiments, only the laser is modulated. For PADMR experiments, only the RF is modulated. See SI section S1.2 for details. (b) The molecular reference frame commonly used for acenes is assumed valid here.33 (c) The unit-cell for crystallites of TSPS-PDT, first reported in ref (10), is reproduced here. The reduced symmetry of TSPS-PDT, relative to standard acenes, means that there are four unique nearest-neighbor-pair conformations, two of which are shown. To aid in visualization, the trialkylsilyl substituents are not included.
PMC10009807
jz2c03665_0001.jpg
0.424045
9e1c3d7907c5428c840663b40e9f317a
Steady-state and photoinduced absorption spectra for a TSPS-PDT film. (a) Absorption spectrum (−log(T/T0)) measured at room temperature with the experimental setup described in the text. A constant value (0.25) was subtracted to correct for scattering. The gap in the spectrum is where the notch filter absorbs. (b) The PA spectrum (642 nm pump wavelength, pump intensity ≈ 2 W/cm2, fmodopt = 1 kHz, temperature ≈ 5–10 K) shows strong absorption peaking at 595 nm, associated with triplet states. The modulation frequency, fmodopt = 1 kHz, is low, so the spectrum approximately represents steady-state PA. This spectrum has been rephased by ϕ = +22.9° to put most of the signal into a single channel (the rephasing process is discussed in section S1.2.5). Inset: PA signals in the NIR are also clear. Experimental parameters for both data sets are the same, except that an 800 nm long-pass filter was added for the NIR data. This spectrum has been rephased by ϕ = −70°.
PMC10009807
jz2c03665_0002.jpg
0.474892
3def8e3196f04eb3af7b560abe36931b
The PADMR spectrum of TSPS-PDT is dominated by triplet signatures. (a) Detecting at the triplet PA’s peak, the f-PADMR spectrum of a TSPS-PDT film shows three main peaks (85, 990, and 1077 MHz), corresponding to transitions within the triplet manifold, T1 (600 nm probe, 642 nm continuous pump, pump intensity ≈ 15 W/cm2, 2.99 kHz RF-modulation frequency, zero magnetic field, temperature ≈ 5–10 K). The negative signals indicate that PA intensity is reduced by RF drive. These data have been rephased by ϕ = +29.8° to maximize the amplitude in a single channel (the rephasing process is discussed in SI sections S1.2.5 and S1.2.6). (b) Sublevel energy diagram for T1 (left) and 5TT (right) in the zero-field basis. Arrows show observed transitions. The T1 splittings are from a fit, with |DT| = 1033 MHz and |ET| = 41.6 MHz. The 5TT sublevels are named as in ref (4), and the splittings shown disregard anisotropic exchange.34 (c) Much weaker features are assigned to quintet transitions.
PMC10009807
jz2c03665_0003.jpg
0.443416
631f8091ff69481a807452f216b8f185
f × λ-PADMR correlates near-IR PA with T1 MR. (a) The PADMR signal magnitude, , shows strong signals when probing the 1000 nm PA (Figure 2) and driving the TZ ↔ TY transition (642 nm pump wavelength, pump intensity ≈ 15 W/cm2, fmodRF = 2.99 kHz, temperature ≈ 5–10 K). Slight asymmetry shows that the PA and MR spectral positions are correlated. (b) λ-PADMR spectra demonstrate shifting PA intensity as a function of fDriveRF. They are slices taken from a global fit of the phase-optimized PADMR data (applied phase, ϕ = +34.2°; see SI section S1.2.5 for a discussion of the rephasing process). (c) The λ-PADMR basis spectra obtained by the fit. All are most intense near 1000 nm. A superimposed positive feature is seen at 850 nm when driven by 972 MHz RF and likely corresponds to a GSB (RF increases ground-state population). It shifts to bluer wavelengths with increasing fDriveRF. A similar shift is seen in a GSB near 740 nm, and pronounced shifts are seen when probing near the 600 nm PA (SI section S2.3.1).
PMC10009807
jz2c03665_0004.jpg
0.465176
26b1852819a64809bedbdbc5cdaa10cf
Triplet-pair model for low-temperature TSPS-PDT films. (a) Dynamical model showing triplet-pair formation (SF), diffusion, and deactivation via singlet-channel and triplet-channel annihilation (SCA and TCA), T1 deactivation through intersystem crossing (ISC), and 1TT ⇌ 5TT equilibration (energy gaps are not to scale). Due to a large antiferromagnetic JTT,105TT forms in low yields. (b) 1TT dissociation yields either TX + TX, TY + TY, or TZ + TZ. X, Y, and Z correspond to TX, TY, and TZ. RF absorption drives one partner into a different sublevel, preventing reassociation to 1TT. (c) An example pathway in which geminate triplet pairs produce a PADMR signal. Molecular stacks are extended along the crystal’s a-axis. 1TT formation is followed by a hop. Spatial separation electronically decouples the triplets (JTT → 0), but they remain spin-correlated: here, as TZ + TZ. Decoupled triplets can absorb RF photons: here, RF with frequency DT – ET drives a TZ → TY transition. With reassociation, strong electronic coupling returns; the new pair configuration has net-triplet or net-quintet character. Both 3TT and 5TT may dissociate again, but 3TT can also deactivate via TCA.
PMC10009807
jz2c03665_0005.jpg
0.450119
9c5acb18677443dbad66041847086604
PRISMA Flow Diagram for Selection Eligible Study
PMC10009940
12889_2023_15321_Fig1_HTML.jpg
0.426146
da29cd7ee7b54a238de75e2d269244d7
Oviposition distribution, natural variation, and the relationship between plant size in oviposition preference by P. rapae butterflies on Arabidopsis accessions. (A) Heatmap showing the average number of eggs deposited per plant in the respective plots over seven independent experiments. The position of the cage is indicated by the compass. (B) Normalized average number of eggs deposited by P. rapae butterflies on 350 different Arabidopsis accessions. Data are the average of seven independent experiments on the total set of 350 accessions, each experiment containing one randomly positioned plant per accession. In each experiment, 10–15 female P. rapae butterflies were allowed to freely oviposit for 2–3 d on the offered population of 350 plants. Error bars show standard errors (±SE). In the color gradient on the right, specific accessions with distinct normalized average egg counts are highlighted. (C) Normalized average number of eggs deposited per plant category (n=4) on small (S), medium (M), and large (L) plants with standard error (±SE) bars. Significance was calculated using Student’s t-test (***P≤0.001 and ****P≤0.0001). (D) Examples of 4-week-old Arabidopsis accessions in the plant size categories.
PMC10010613
erac501f0001.jpg
0.420534
a7fff572f60a4ff8981bc6a7d3cbdaca
GWAS and fine mapping results for oviposition preference of P. rapae on 346 Arabidopsis accessions. (A) Manhattan plot (grey and dark grey) showing the –log10(p) values of the SNP–trait associations from the GWAS results on Arabidopsis chromosomes 1–5 (x-axis). Narrow sense heritability was estimated to be very low (h2=0.0014 with a 95% confidence interval of 0.0–1.0; Kruijer et al., 2015). The blue line indicates the arbitrary LOD threshold of 4.0 (–log10(p)=4.0) for selection of SNPs. Above the Manhattan plot a gene cluster is depicted (http://signal.salk.edu/atg1001/3.0/gebrowser.php) that was found upstream of the transposable element gene AT3G25725 in which SNPs above the threshold were found by the GWAS. In bold three GWAS loci are indicated for which SNP–trait associations (LOD≥4.0) were confirmed via fine mapping. LTR, long terminal repeat. (B) Fine mapping (FM; grey dots) of three SNP–trait associations that were identified by the GWAS (green dots), using the 50-kb window around the GWAS SNPs from the genome sequences of 164 of the tested Arabidopsis accessions. The graphs show the –log10(p) values of the SNP–trait associations on the y-axis and the chromosome position of the SNPs in base pairs (bp) on the x-axis. Significant (FDR-corrected) FM associations are shown in black ATG numbers along with the number of significant associations in the FM summary.
PMC10010613
erac501f0002.jpg
0.384467
430c916a033c442493df16d215f1b901
Amino-acid changes confirm possible altered gene function of fine mapping candidates. Manhattan plots of fine mapping results (MAF>5%, FDR-corrected) for the candidate genes AT3G25760 and AT4G04450, including both introns and exons or only exons. y-Axes show the –log10(p) and the x-axes the position in base pairs (bp). For each gene a model of the introns and exons is shown according to the 1001 genomes browser (http://signal.salk.edu/atg1001/3.0/gebrowser.php). A zigzag indicates missing sequencing data (AT3G25760). Important (significant) nucleotides are indicated with colors: A (red), T (ochre/yellow), C (blue), and G (green). Non-synonymous amino acid changes are depicted with red letters.
PMC10010613
erac501f0003.jpg
0.405448
83ee0adecb224d789baba1c17380aa74
Oviposition choice assay on fine-mapping confirmed genes. Six mutants, including a MYC-triple mutant (myc234) that lacks glucosinolates and a JA lacking mutant (aos) control, a T-DNA insertion line (wrky42) and an RNAi line (AOC::RNAi line 16-1) were tested for oviposition preference by P. brassicae in a two-plant choice assay. Replicates of each two-plant assay are indicated as n (n=15–22). Bars represent the average distribution of eggs between wild-type Col-0 plants (dark grey bars) and the mutant plants (light grey bars) in a choice test. Significant differences were calculated using Student’s t-test (non-significant, ns; *P≤0.05, **P≤0.01, ***P≤0.001, and ****P≤0.0001).
PMC10010613
erac501f0004.jpg
0.487239
2e0d905ffeb543aeaca907620025aded
Caterpillar performance on mutant wrky42. Graph showing the average caterpillar weight in grams along with standard error (±SE) of mean error bars in a no-choice test. Caterpillar weight was measured after placing one L1 P. brassicae caterpillar on either a mature wild-type Col-0 plant (n=13) or a wrky42 mutant plant (n=14), from which deposited eggs were removed prior to caterpillar exposure, and allowing them to feed for 13 d. Plant replacements were added well before food was becoming scarce. Significant differences were calculated using Student’s t-test (****P≤0.0001).
PMC10010613
erac501f0005.jpg
0.446533
b5c3dcc994da4005b35d94a41b9f981d
Axial magnetic resonance imaging revealing cerebellar and pontine volume loss with crossed hyperintensity of the pons, known as “hot cross bun sign.”
PMC10010663
ajpmr-102-e54-g001.jpg
0.465757
cc5e280ac30447388db3b7de12e2b265
Distribution of aldosterone synthase (CYP11B2) and LHCGR in the patient’s adrenal. (A), Distribution of CYP11B2 immunoreactivity in the tumor region (T) of the adrenal tissue. (B-E), Distribution of CYP11B2 (B, D) and LHCGR (C, E) immunoreactivities in consecutive sections of the APA tissue at low (B, C) and high (D, E) magnifications. Similar distribution of CYP11B2 and LHCGR immunoreactivities were observed in some areas (arrows) in (B, C). PT indicates peritumoral tissue; and V, vein.
PMC10011616
fendo-14-1153374-g001.jpg
0.417997
2e2e5ecfea504ce9b2937a65d0d8fd9d
Diagram 1. Overview of shared pathophysiology and differential outcomes of DM and frailty.
PMC10012063
fcdhc-03-895313-g001.jpg
0.392325
8b08839f7cfe4f7a90a8cb18ba6c7f8e
Histogram of all CGM measurements. Abbreviation: CGM, continuous glucose monitoring.
PMC10012361
10.1177_19322968211059547-fig1.jpg
0.435041
8f35fe07c3e4476b802922fcf7b7847e
Pairwise scatterplots, distribution and linear correlation coefficients of the investigated metrics. Distribution of each metric can be found in the main diagonal, pairwise correlation coefficients in the upper right triangle and their pairwise scatterplots in the bottom left half. The 9 highest risk patients according to the estimated time above 600 mg/dl obtained with EVS are highlighted. Abbreviations: CONGA, continuous net overall glycemic action; CV, coefficient of variation; EVS, extreme value statistics; Hrs 400+, EVS estimation of hours spent above 400 mg/dl per year; Hrs 600+, EVS estimation of hours spent above 600 mg/dl per year; IQR, interquartile range; MAGE, mean amplitude of glycemic excursions; RL, return level; TAR (>250), observed time above range (TAR) spent above >250 mg/dl.
PMC10012361
10.1177_19322968211059547-fig2.jpg
0.440346
059dc01dbeee425bb2b6ca10de8cbad3
Histograms of CGM measurements of the patients with the highest risk. Abbreviation: CGM, continuous glucose monitoring.
PMC10012361
10.1177_19322968211059547-fig3.jpg
0.429115
f9a5a8e25a5444ec9c89b027ef41d7af
The effect of BMI on the distribution of the hourly maximum blood glucose. Abbreviation: BMI, Body Mass Index.
PMC10012361
10.1177_19322968211059547-fig4.jpg
0.428143
6ae663f327d0486d8297d451b1eacebb
Impact of upper saturation level (trimming) on traditional GV metrics and metrics based on EVS. Values obtained with the given upper saturation level were divided with the original value (i.e., when only the physical 400 mg/dl saturation level was in effect). Abbreviations: CONGA, continuous net overall glycemic action; CV, coefficient of variation; EVS, extreme value statistics; GV, glycemic variability; Hrs 400+, EVS estimation of hours spent above 400 mg/dl per year; Hrs 600+, EVS estimation of hours spent above 600 mg/dl per year; IQR, interquartile range; MAGE, mean amplitude of glycemic excursions; RL, return level; TAR (>250), observed time above range (TAR) spent above >250 mg/dl.
PMC10012361
10.1177_19322968211059547-fig5.jpg
0.485342
857420b3baa3457ba31ac2df46a01b66
Kaplan-Meier curves for A, local failure-free survival; B, distant metastasis-free survival; C, progression-free survival; and D, overall survival in patients in the retrospective clinical cohort with an on-treatment mANC ≤ 5,000 cells/µL vs mANC > 5,000 cells/uL. HR, hazard ratio; mANC, mean absolute neutrophil count.
PMC10014332
gr1.jpg
0.371656
d3b6d05ee3c547e6afdae0d5547f3af8
Circulating and tumor-associated myeloid cells during Chemoradiotherapy. A and B, CD11b+CD11c- cells in PBMC (A) and tumor cytobrush (B) samples. C, Scatterplot showing a correlation between cytobrush and PBMC CD11b+CD11c- samples. The black regression line formula and R2 refer to the overall population of samples, while the green regression line formula and R2 refer to the regression line for T4 samples only (line not shown). PBMC, peripheral blood mononuclear cells.
PMC10014332
gr2.jpg
0.464479
f75773088eaf4da49fad8e893d836024
Flow diagram for the identification and exclusion of studies. MNI, Montreal Neurological Institute; DMN, default mode network.
PMC10014826
fnins-17-1136790-g001.jpg
0.420611
0473b8d323f74bd881f2a4af4bfad5ba
The areas of increased (red) and decreased (blue) functional connectivity (FC) in the main meta-analysis. “R” and “L” denote the right and left sides of the brain, respectively. The color bar indicates the maximum and minimum seed-based d mapping (SDM)-Z value.
PMC10014826
fnins-17-1136790-g002.jpg
0.451328
7116d1ffd9224968911a21ce8aadcc3c
Initial subcutaneous calcium deposits along the length of right arm. At initial development, these findings followed the same path of the catheter used for parenteral nutrition and as such, a local reaction to an extravasation was firstly hypothesized.
PMC10015177
gr1.jpg
0.42364
c389e3c9757845bdb02def67ffe8c76e
Evolution of the calcifications along the right arm 4 months from the previous photograph. The lesion became more extensive both clinically and radiologically, prompting histological and genetic analysis. No surgical management had been performed by this stage.
PMC10015177
gr2.jpg
0.467684
6ad1847e9990477bab8a6523643b75f3
Map of the survey sites, in their respective maize agroecological zones.
PMC10015270
gr1.jpg
0.431466
b1d68652785d4894a6b2a1a923c1bc18
Trends in the adoption of improved maize varieties (IMV) in Kenya from 1992 to 2013, in adoption rate (% of farmers adopting IMVs) and adoption intensity (% of maize area in IMVs) (error bars represent standard errors).
PMC10015270
gr2.jpg
0.380338
360621efb8e34e70acceabb58cf2ce79
Area share of maize varieties, by source and over time.
PMC10015270
gr3.jpg
0.502314
c047ea83c99a4722ae7e8429849509cf
Area share of different types of maize varieties by source and agroecological zone, over time.
PMC10015270
gr4.jpg
0.451111
1880990eab114f84943de38a944b578a
Trends in weighted average age (years) of maize varieties, for all varieties and for improved varieties only.
PMC10015270
gr5.jpg
0.462601
19cc1e47b49b4bde84200671a59afa02
Trends in weighted average age (WAA, in years) of maize varieties, all varieties and improved varieties alone, by agroecological zone.
PMC10015270
gr6.jpg
0.45515
66b78e48153c4c5ca3008a791f73224b
Scalp psoriasis status measured using the Scalp Physician Global Assessment (Scalp-PGA) at baseline and after 4 weeks' treatment with Cal/BD aerosol foam (n = 193); Scalp-PGA was assessed with the three categories “mild/moderate/severe” [5], with the category “mild” being the lowest possible category to rank, thus also including “clear” lesions.
PMC10015748
drm-0239-0206-g01.jpg
0.413156
08171fdd26324c6994cd69caa566cd42
Severity of scalp psoriasis during the course of the 4-week study measured using the scalp psoriasis patient global assessment (Scalp-PaGA, 0 = clear, 5 = very severe; n = 193; LOCF; mean ± SD); *p < 0.001 versus baseline.
PMC10015748
drm-0239-0206-g02.jpg
0.379295
31ccd01852ba4a00af55ec1cedde7bda
Severity of scalp psoriasis during the course of the 4-week study measured using the scalp psoriasis patient global assessment (Scalp-PaGA) (n = 193).
PMC10015748
drm-0239-0206-g03.jpg
0.381592
816971f419a64c6f815bda0b646ecf20
Patient reported perception of itch on the affected scalp rated using a NRS ranging from 0 = no itch to 10 = maximum itch at baseline and after 4 weeks of treatment with Cal/BD aerosol foam (n = 194).
PMC10015748
drm-0239-0206-g04.jpg
0.409992
36b4a95ea5ca482198b847bad8069745
Influence of scalp psoriasis on the QoL of patients measured using Dermatology Quality of Life Index (DLQI) at the start of therapy and end of study following 4 weeks' treatment with Cal/BD aerosol foam (n = 186).
PMC10015748
drm-0239-0206-g05.jpg
0.459242
34dea35184004fb7807d6e2d40d79822
Patient reported perception of Cal/BD aerosol foam feeling on the skin at day 3 (each category rated using a 5-point scale ranging from “very strong” to “not at all”).
PMC10015748
drm-0239-0206-g06.jpg
0.45185
71087fe96f7542feaca62d5cc13e9404
Hypoxia-associated circRNA profiling and expression characteristics of Hsa_circ_0000566 in osteosarcoma (OS). (A) CircRNA microarray analysis reveals 35 upregulated and 23 downregulated circRNAs in OS cells under normoxic and hypoxic conditions. The black arrow represents Hsa_circ_0000566. (B) OS cells incubated under various oxygen concentrations. Total RNA extraction was performed for qRT-PCR assay. Western blotting was performed to determine the protein level of HIF-1α. Results are reported as mean ± standard deviation (SD), *p < 0.05, n = 3. Scale bars, 200 μm. (C) Hsa_circ_0000566 expression is much higher in primary OS tissue than in chondroma tissue. Results are representative images according to three different experiments. (D) Quantitative real-time polymerase chain reaction (qRT-PCR) results comparing Hsa_circ_0000566 mRNA expression in 12 OS and chondroma samples. Results are reported as mean ± SD, *p < 0.05, n = 12. (E) Hsa_circ_0000566 expression levels in hFOB1.19 and various OS cell lines. Results are reported as mean ± SD, *p < 0.05, n = 3. (F) Schematic diagram showing Hsa_circ_0000566 back-spliced by exons 2-11 of the VRK1 gene and the corresponding Sanger sequencing. (G) RT-PCR results validating the presence of Hsa_circ_0000566 in 143B and HOS cells. Various primers amplified the Hsa_circ_0000566 region in cDNA but not in genomic DNA. β-actin was used as the negative control. Divergent primers are presented as the opposite direction of the arrowhead, and the convergent primers were shown as the face-to-face direction of the arrowhead. (H) RT-PCR results indicating Hsa_circ_0000566 and VRK1 mRNA expression in untreated 143B and HOS cells and in the cells subjected to treatment with RNase-R. (I) RNA fluorescence in situ hybridization (FISH) results revealing Hsa_circ_0000566 localized mainly in the cytoplasm. Hsa_circ_0000566 probes were labeled with cy3 and nuclei were stained with 4’,6-diamidino-2-phenylindole (DAPI). Scale bars, 100 μm. (J) qRT-PCR determination of the main localization of Hsa_circ_0000566 in OS cells. Results are reported as mean ± SD, *p < 0.05, n = 3.
PMC10017158
AD-14-2-529-g1.jpg
0.441668
0b713e2ebd174309b9896e5dab0dbbc1
Hsa_circ_0000566 contributes to in vitro osteosarcoma (OS) cell progression under hypoxic conditions. (A) Hsa_circ_0000566 overexpression and knockdown induced and repressed OS cell proliferation under hypoxia. Results are reported as mean ± standard deviation (SD), *p < 0.05, n = 3. Circ_0000566 represents Hsa_circ_0000566 overexpression, and si circ_0000566 represents Hsa_circ_0000566 knockdown. Vector and Si NC represents the negative control of Hsa_circ_0000566 overexpression and Hsa_circ_0000566 knockdown, respectively. (B) EdU exhibits the impact of Hsa_circ_0000566 on OS cell proliferation under hypoxia. Nuclei are stained with 4’,6-diamidino-2-phenylindole (DAPI). Results are reported as mean ± SD, *p < 0.05, n = 3. Scale bars, 100 μm. (C) Colony formation experiment verifies Hsa_circ_0000566 functions in OS cells under hypoxia. Results are reported as mean ± SD, *p < 0.05, n = 3. (D) Soft agar colony formation assay indicates the effects of Hsa_circ_0000566 on 143B and HOS cell colony forming capacity under hypoxia. Results are reported as mean ± SD, *p < 0.05, n = 3. Scale bars, 100 μm. (E) OS cell migration capacity as determined by Transwell™ migration assays. Results are reported as mean ± SD, *p < 0.05, n = 3. Scale bars, 100 μm. (F) Flow cytometry verifies Hsa_circ_0000566 functions in OS cell apoptosis. Results are reported as mean ± SD, *p < 0.05, n = 3.
PMC10017158
AD-14-2-529-g2.jpg
0.457455
9f338c314e5641cabf9cdea6cfa20577
Hsa_circ_0000566 accelerates osteosarcoma (OS) glucose metabolism and regulates hypoxia-enhanced glycolysis. (A) Colors of the media indicate that Hsa_circ_0000566 silencing decreased lactate accumulation under hypoxia. (B-C) Quantitative real-time polymerase chain reaction (qRT-PCR) or western blots evaluating the expression levels of genes involved in glucose metabolism in 143B and HOS cells transfected with Hsa_circ_0000566-overexpressing, Hsa_circ_0000566 (shRNA), or vector plasmids. Results are reported as mean ± standard deviation (SD), *p < 0.05, n = 3. (D) Hsa_circ_0000566 knockdown in OS cells with decreased lactate accumulation, while Hsa_circ_0000566 overexpression has increased lactate accumulation. Results are reported as mean ± SD, *p < 0.05, n = 3. (E) Extracellular acidification rate (ECAR) indicates glycolysis rate. ECAR decreases in response to Hsa_circ_0000566 knockdown and increases in response to Hsa_circ_0000566 overexpression. Oxygen consumption rate (OCR) represented mitochondrial respiratory capacity. OCR is enhanced in response to Hsa_circ_0000566 silencing and reduced in response to Hsa_circ_0000566 overexpression in OS cells. Results are reported as mean ± SD, *p < 0.05, n = 3.
PMC10017158
AD-14-2-529-g3.jpg
0.480262
19898175fbe842adb63973a82fd3e667
Hsa_circ_0000566 establishes interactions with HIF-1α and confers protection against ubiquitination-mediating degradation. (A) Effects of Hsa_circ_0000566 knockdown and Hsa_circ_0000566 overexpression on mRNA and protein expression in 143B and HOS cells under hypoxia. Results are reported as mean ± standard deviation (SD), *p < 0.05, n = 3. (B) Western blotting results revealing the impact of bortezomib treatment on the changes occurring at HIF-1α protein level mediated by Hsa_circ_0000566 silencing and vector transfection. (C) Western blotting assessment of the impact of CHX treatment on the variations in HIF-1α protein levels affected by Hsa_circ_0000566 silencing and vectors. Results are reported as mean ± SD, *p < 0.05, n = 3. (D) The western blot illustrates the effects of Hsa_circ_0000566 knockdown in the Hyp564 HIF-1α protein levels in the presence or absence of bortezomib treatment. (E) Immunoprecipitation assessing the HIF-1α ubiquitination levels in Hsa_circ_0000566 silencing and Hsa_circ_0000566 overexpressing osteosarcoma (OS) cells under hypoxia. Culture media were supplemented with bortezomib (250 nM) for 6 h. (F) The combination of Hsa_circ_0000566 with HIF-1α confirmed by radioimmunoprecipitation (RIP). Results are reported as mean ± SD, *p < 0.05, n = 3. (G) Pulldown assay validation of the interaction between Hsa_circ_0000566 and HIF-1α. (H) A RIP assay of HIF-1α regions interacting with Hsa_circ_0000566. Schematic diagram shows HIF-1α protein fragments. Results are reported as mean ± SD, *p < 0.05, n = 3. (I) Interaction profile between Hsa_circ_0000566 and HIF-1α obtained from catRAPID (left). (J) Schematic diagram showing Hsa_circ_0000566 RNA fragments. Combinative regions between Hsa_circ_0000566 and HIF-1α were identified by RIP assay. Results are reported as mean ± SD, *p < 0.05, n = 3.
PMC10017158
AD-14-2-529-g4.jpg
0.438636
39d8698940dd48428b3959179825d5bf
Hypoxia-induced Hsa_circ_0000566 stabilizes HIF-1α by attenuating interactions between Von Hippel—Lindau (VHL) and HIF-1α. (A) The interaction between VHL and Hsa_circ_0000566, assessed by performing a RIP assay. Results are reported as mean ± standard deviation (SD), *p < 0.05, n = 3. (B) Interaction between VHL and Hsa_circ_0000566, identified by conducting pulldown assays. (C) Effects of Hsa_circ_0000566 silencing and Hsa_circ_0000566 overexpression on VHL protein and mRNA expression, identified by performing western blotting and quantitative real-time polymerase chain reaction (qRT-PCR). (D) Immunoprecipitation experiment revealing VHL-HIF-1α interaction that is inhibited by Hsa_circ_0000566 overexpression. (E) Sequential coimmunoprecipitation assay, performed to determine whether Hsa_circ_0000566 could simultaneously bind VHL and HIF-1α.
PMC10017158
AD-14-2-529-g5.jpg
0.41502
d22fd747ab9c47e6a4d5df5cea71527c
HIF-1α overexpression reverses Hsa_circ_0000566 silencing-induced attenuation of osteosarcoma (OS) cell proliferation, migration, and glucose metabolism. (A) Hsa_circ_0000566 silencing in OS cells inhibits LDHA enzyme activity, whereas HIF-1α overexpression promotes enzyme activity. (B-C) Quantitative real-time polymerase chain reaction (qRT-PCR) and western blots identifying the effects of HIF-1α overexpression in Hsa_circ_0000566 knockdown OS cells. Results are reported as mean ± standard deviation (SD), *p < 0.05, n = 3. (D) A luciferase report assay showing that Hsa_circ_0000566 knockdown markedly reversed HIF-1α transcription induced by hypoxic stress. Results are reported as mean ± SD, *p < 0.05, n = 3. (E-F) HIF-1α overexpression recovered Hsa_circ_0000566 knockdown-induced decreases in glucose uptake and lactate production. Results are reported as mean ± SD, *p < 0.05, n = 3. (G) Extracellular acidification rate (ECAR) and oxygen consumption rate (OCR) indicate that HIF-1α overexpression recovered the decline in glycolysis rate induced by Hsa_circ_0000566 knockdown under hypoxia. Results are reported as mean ± SD, *p < 0.05, n = 3. CCK-8 experiments reveal that HIF-1α and Hsa_circ_0000566 silencing affects OS cell proliferation under hypoxia. Results are reported as mean ± SD, *p < 0.05, n = 3. (I) EdU assay shows that HIF-1α and Hsa_circ_0000566 silencing influences OS cell vitality. Scale bars, 100 μm. (J) Colony formation assay indicates that colony formation ability is mediated by HIF-1α and Hsa_circ_0000566 knockdown. Results are reported as mean ± SD, *p < 0.05, n = 3. (K) Effects of HIF-1α and Hsa_circ_0000566 attenuation on tumor migration, as evidenced by Transwell™ migration assay results. Results are reported as mean ± SD, *p < 0.05, n = 3. Scale bars, 100 μm. (L) Flow cytometry evaluates the impact of HIF-1α and Hsa_circ_0000566 attenuation on OS cell apoptosis. Results are reported as mean ± SD, *p < 0.05, n = 3.
PMC10017158
AD-14-2-529-g6.jpg
0.477065
d44444456f3d4f9a8764cf5638430e68
HIF-1α targets and regulates Hsa_circ_0000566 in glycolysis under hypoxic stress. (A-B) Quantitative real-time polymerase chain reaction (qRT-PCR) and western blotting demonstrates that among vital transcription factors such as HIF-1α, HIF-2α, p53, and Hsa_circ_0000566, only Hsa_circ_0000566 is induced by HIF-1α under hypoxia. Results are reported as mean ± standard deviation (SD), *p < 0.05, n = 3. (C) The HIF-1 motif. (D) Luciferase report assays identifying the effective binding regions on Hsa_circ_0000566. Results are reported as mean ± SD, *p < 0.05, n = 3. (E) Luciferase gene experiment shows that mutant 1 is the binding site that regulates Hsa_circ_0000566 transcription. Results are reported as mean ± SD, *p < 0.05, n = 3. (F-H) ChIP assay indicates that HIF-1α is not bound to the genomic region after mutation of HRE2 mutant 1. Results are reported as mean ± SD, *p < 0.05, n = 3. (I) Schematic diagram of the HIF-1α/Hsa_circ_0000566/HIF-1α loop.
PMC10017158
AD-14-2-529-g7.jpg
0.392291
73e6ccf672ca4fa1b0aeb37f1747e318
Hsa_circ_0000566 promotes osteosarcoma (OS) glucose metabolism and tumorigenesis progression in vivo. (A) 143B cells stably transfected with Hsa_circ_0000566 knockdown, HIF-1α overexpression, or empty vector plasmids. Nude mice were subcutaneously injected with 1 × 107 cells that were either stable negative controls or those with Hsa_circ_0000566 knockdown, HIF-1α overexpression, or Hsa_circ_0000566 knockdown. Thirty days after injection, the animals were euthanized, and their tumors dissected and photographed. (B) Tumor weight measurements on the same day the mice were euthanized. Results are reported as mean ± standard deviation (SD), *p < 0.05, n = 5. (C) Tumor volumes (ab2/2) were calculated every 6 d from the day after the mice were injected with stable OS cells. (D-E) Western blotting and quantitative real-time polymerase chain reaction (qRT-PCR) exhibit the expression levels of the genes involved in glycolysis metabolism. Results are reported as mean ± SD, *p < 0.05, n = 3. (F) Fluorescence in situ hybridization (FISH), hematoxylin and eosin (H&E) staining, and immunohistochemistry (IHC) analysis indicate the OS organization in mice and relative GLUT1, GLUT4, PDK1, PDK4, and LDHA protein levels in tumors from different groups. (G) In situ tumor formation experiment reveals that HIF-1α overexpression recovered Hsa_circ_0000566 knockdown-induced tumor attenuation. Results are reported as mean ± SD, *p < 0.05, n = 4. (H) Micro-computed tomography (CT) indicates the functions of HIF-1α and Hsa_circ_0000566 knockdown in bone loss. (I) H&E staining of lung metastasis. In mice injected in the tail vein with various stable 143B cells, lung metastasis was detected using an in vivo bioluminescence imaging system. Results are reported as mean ± SD, *p < 0.05, n = 5.
PMC10017158
AD-14-2-529-g8.jpg
0.43418
bccd2fcddfb8419c826352ad7867f5be
Characteristics of three-dimensional fluorescence spectra.
PMC10017222
CIN2023-7535594.001.jpg
0.419517
c794dc395d6143148b2dd73154bbe641
The neuronal structure.
PMC10017222
CIN2023-7535594.002.jpg
0.445836
d67cc27b03fe4299b1a8ff49a71cc9f3
The deep learning model.
PMC10017222
CIN2023-7535594.003.jpg
0.4195
ca784e66025b420584d8078e4cd88cf8
Main properties of biochar.
PMC10017222
CIN2023-7535594.004.jpg
0.464254
07947efe72d543a98489c1771e59810e
Effects of different chemical fertilizers on the content of water-soluble organic matter in the soil at 0∼60 cm depth.
PMC10017222
CIN2023-7535594.005.jpg
0.461193
059c840a82564818bc788b36a6fdb799
Three-dimensional fluorescence spectra of water-soluble organic matter after 6 years of fertilization (I tyrosine-like substance; II: tryptophan-like substance; III: fulvic acid-like substance; IV: soluble microbial by-products; V: humic acid-like substance): (a) no fertilizer; (b) NPK; (c) BC; (d) NPK + BC; (e) N + BC.
PMC10017222
CIN2023-7535594.006.jpg
0.470748
0538c6b524764561acff55bc6322d1d6
Results of different fluorescent components (C1: humic acid-like substance; C2: fulvic acid-like substance; C3: protein-like substance): (a) CK; (b) NPK; (c) BC; (d) NPK + BC; (e) N + BC.
PMC10017222
CIN2023-7535594.007.jpg
0.475009
89e8059dd7924795be57aeb5ffb3166d
Different FIs in the whole soil depth:(a) FI; (b) HIX; (c) BIX. T0: no fertilizer; T1: NPK; T2: BC; T3: BC + NPK; T4: BC + N.
PMC10017222
CIN2023-7535594.008.jpg
0.411436
f4ad1be2d0654ca5a0d425e8038ade46
Pericytes and mesenchymal stromal cells in vivo and in vitro. The upper left portion of the figure shows a schematic representation of a portion of a tissue, where a small blood vessel formed by endothelial cells is surrounded by vascular smooth muscle cells at the arteriole level, and by different types of pericytes, including a transitional form that represents an intermediate between vascular smooth muscle cells and capillary pericytes. A paravascular interstitial cell is also represented, as well as a monocyte, and tissue-specific cells. The lower left portion of this figure represents the fact that not only pericytes, but various cell types including some mature tissue-specific cells, endothelial cells, and even monocytes have been found to be able to give rise to cells bearing mesenchymal stromal cell characteristics in culture. The right half of the figure schematically depicts the conversion of monocytes into M1 macrophages after a tissue injury event, in the inflammatory stage of wound healing (for sake of simplification, tissue-resident macrophages are not depicted). These M1 macrophages secrete pro-inflammatory factors, which lead to pericyte activation and, possibly, paravascular interstitial cell proliferation. In the proliferative stage of wound healing, the progeny of perivascular cells secrete molecules that contribute to change the balance between pro-inflammatory and pro-regenerative macrophages toward an increase of the latter type, referred to as “macrophages of the M2 spectrum”. As these reparative macrophages become more prevalent, they produce increased numbers of anti-inflammatory and pro-regenerative molecules. In the reparative stage of the wound healing process, reparative macrophages contribute to angiogenesis, which restores blood supply to the tissue; additionally, some of the previously activated pericytes may re-acquire a resting pericyte phenotype to provide physical and functional support to the newly formed blood vessels. Soluble molecules produced by cultured mesenchymal stromal cells (whether in extracellular vesicles or not), or cultured mesenchymal stromal cells themselves, can be used to favor the acquisition of a reparative phenotype by monocytes and M1 macrophages, as represented by the arrow that connects the lower left and right portions of the figure.
PMC10017474
fcell-11-1148121-g001.jpg
0.510297
5705642011e246cc9775f25738c81cf4
A SEM images of the synthesized organosilica particles PMO-OH at 50k of magnification (inset: 200k of magnification). B Size distribution by volume (H2O, 25°C) of PMO-OH nanoparticles measured by hydrodynamic light scattering (DLS)
PMC10017643
604_2023_5679_Fig1_HTML.jpg
0.427607
8a00bb47fbed460c93c24de9ac0ebd65
SEM images of the reactive side of POM@PMO-PE sensor before phosphate detection using different scales and orders of magnitudes A ×250, B ×5k, and C ×50k showing the embedment of PMOs nanoparticles in between carbon graphite micro-flakes and D corresponding EDX spectra
PMC10017643
604_2023_5679_Fig2_HTML.jpg
0.488366
cc53815a4f754b12821f8dcb498c4904
Comparison between the anodic SWV response POM-PE and POM@PMO-PE in 1M H2SO4 prepared in KCl 0.1M (black-dashed curves) and after adding 10nM of phosphate as NaH2PO4 (solid curves)
PMC10017643
604_2023_5679_Fig3_HTML.jpg
0.400155
6d2dee506bab485e98a9f4738e59ca19
A Anodic SWV response of activated POM@PMO-PE sensors upon successive additions of aqueous H2PO4− standards prepared in 1M H2SO4/0.1M KCl (blank) to obtain increasing concentrations 1–500nM. B Calibration curves were obtained based on the relative variation of the maximum peak intensity corrected by subtracting any blank noise (Ip–I0) at the corresponding potentials (−0.06 to −0.05V) as a function of cumulative phosphate concentrations (n=3)
PMC10017643
604_2023_5679_Fig4_HTML.jpg
0.443837
a70a5f9600a1418f98a1b92cca5b7659
Anodic SWV response of activated POM@PMO-PE sensors toward two different concentrations of phosphate and silicate as a potential interferent. The error bars are the RSD of three independent measurements (n=3)
PMC10017643
604_2023_5679_Fig5_HTML.jpg
0.429852
8487803bc32747428fc720f2bd801990
Experimental procedure for periodic mesoporous organosilica (PMO) preparation starting from CTAB [Created with BioRender.com].
PMC10017643
604_2023_5679_Sch1_HTML.jpg
0.420724
77a1520775444b9cb7a30a358df8c678
Flow diagram of cohort study population
PMC10017785
464_2022_9670_Fig1_HTML.jpg
0.403017
194078bac3c64bbab295f9cafcbab497
Signaling network of the IL-2 cytokine family. Receptors of IL-2, IL-15, IL-21, IL-4, IL-7 and IL-9 share a common γ chain subunit. They phosphorylate various STAT proteins downstream by activating JAK/STAT signaling pathway.
PMC10018032
fimmu-14-1090311-g001.jpg
0.451662
1470b34c5f7c48c49a4950ee0205800b
Representative engineering cytokine patterns.
PMC10018032
fimmu-14-1090311-g002.jpg
0.418237
d1e599d155e94327a40dbbdf283a04c7
Patient flow diagram. CGI-I, Clinical Global Impression-Improvement; eCRF, electronic case report form; IRB, institutional review board.
PMC10018127
fpsyt-14-1075939-g001.jpg
0.52226
e11ea66ddd7244a8a281c84c50df3118
Mean weight at baseline, Visit 2 and Visit 3. Error bars represent standard deviations.
PMC10018127
fpsyt-14-1075939-g002.jpg
0.439152
008681b8d9884d8b8fb98cf67cebf698
Change in MADRS total score from baseline. The mean MADRS total score at Visit 1 was based on all patients, not just those that also had data at Visit 2 or Visit 3, however, when analyzing the difference between the visits (Visit 1 to Visit 2 and Visit 1 to Visit 3), scores of patients who had scores at both visits were included. Therefore paired t-tests assessed the difference at Visit 2 (8±2 week) versus Visit 1 (0 week) in 1,890 patients and difference at Visit 3 (24±2 week) versus Visit 1 (0 week) in 577 patients. Error bars represent standard deviations. MADRS, Montgomery-Asberg Depression Rating Scale.
PMC10018127
fpsyt-14-1075939-g003.jpg
0.445159
776807acda8440df935349c99f188bfa
Effectiveness using CGI-I assessment. Effectiveness of vortioxetine was defined as the proportion of patients that showed improvements in CGI-I scores from baseline at Visits 2 (8±2 weeks) and 3 (24±2 weeks). Effectiveness was calculated by classifying “Improved” categories as “effective”, and categories under “No change” and “Worse” as “ineffective”. CGI-I, Clinical Global Impression-Improvement.
PMC10018127
fpsyt-14-1075939-g004.jpg
0.449216
74792b6895e94f008b216b11b5a6e4e9
Change in PDQ-K total score from baseline. The mean PDQ-K total score at Visit 1 was based on all patients, not just those that also had data at Visit 2 or Visit 3, however, when analyzing the difference between the visits (Visit 1 to Visit 2 and Visit 1 to Visit 3), scores of patients who had scores at both visits were included. Therefore paired t-test assessed the difference at Visit 2 (8±2 week) versus Visit 1 (0 week) in 1,595 patients and difference at Visit 3 (24±2 week) versus Visit 1 (0 week) in 476 patients. Error bars represent standard deviations. PDQ-K, Korean version of Perceived Deficits Questionnaire-Depression.
PMC10018127
fpsyt-14-1075939-g005.jpg
0.448698
ee7d842decf64d6c99c44be9aadefe63
Change in DSST total score from baseline. The mean DSST total score at Visit 1 was based on all patients, not just those that also had data at Visit 2 or Visit 3, however, when analyzing the difference between the visits (Visit 1 to Visit 2 and Visit 1 to Visit 3), scores of patients who had scores at both visits were included. Therefore paired t-test assessed the difference at Visit 2 (8±2 week) versus Visit 1 (0 week) in 565 patients or difference at Visit 3 (24±2 week) versus Visit 1 (0 week) in 187 patients. Error bars represent standard deviations. DSST, Digit Symbol Substitution Test.
PMC10018127
fpsyt-14-1075939-g006.jpg
0.506545
7e6119e5f26542d592fdfe26ad346438
Pressure-volume loop demonstrating the phenomenon of “beaking.” A typical pressure-volume loop generated from ventilation of an optimally-expanded lung (solid line) is shown for reference. Note that in the normal condition, a relatively small change in pressure results in a significant change in volume. Excessive positive end expiratory pressure can result in ventilation of a hyperinflated lung (dashed line), and as tidal volume nears total lung capacity it requires a higher change in pressure and a higher peak pressure to attain a similar change in volume. The flattening of the upper portion of the curve is referred to as “beaking” (asterix)
PMC10018229
fped-11-1087857-g001.jpg
0.50015
e89a134ce1884b33b8701ca0c05a8605
Ventilator data flowsheet encompassing 18–21 days of age for a former 27 week, 1,160 gram premature infant with evolving chronic lung disease and supported on pressure-regulated volume control ventilation. As exhaled tidal volume (open diamonds) exceeds the set tidal volume (open circles) (asterix), note that the ventilator compensates by lowering peak inspiratory pressure (closed circles) (arrow).
PMC10018229
fped-11-1087857-g002.jpg
0.438233
1c96185299204cdfb2700477107a0f3d
Chest computed tomography of an 8 month old former 26 week premature infant with severe BPD demonstrating dramatic pulmonary heterogeneity with alveolar simplification of the right upper lobe and diffuse fibrosis of the left lung. The patient is also severely hyperinflated with flattened diaphragms.
PMC10018229
fped-11-1087857-g003.jpg
0.434285
cbeb02f5212c4274ab0d7a0475246c42
Ventilator waveforms from a 2 month old former 24 week premature infant with severe bronchopulmonary dysplasia using two different PEEP strategies. On physical exam, the child has a respiratory rate of 26 breaths/min. A) However, using a PEEP 10 cm H2O, he has multiple failed triggers (yellow circles) and does not appear to breathe above the mandated rate. B) By increasing the PEEP to 17 cm H2O there is improved synchronization with the ability to trigger his spontaneous breaths, though the trigger is delayed (white circles) and one patient effort fails to trigger a breath (yellow circles).
PMC10018229
fped-11-1087857-g004.jpg
0.422995
e9357b33a7e24650a6777fadaef5176c
Ventilator waveforms from an 8 month old former 24 week premature infant with severe bronchopulmonary dysplasia using two different inspiratory time strategies. (A) Using a short (0.4 s) inspiratory time, there is incomplete filling (white arrows) with reduced tidal volume (yellow arrows), increased respiratory rate, and increased peak pressures (red arrows) compared with (B) a long (0.9 s) inspiratory time.
PMC10018229
fped-11-1087857-g005.jpg
0.467656
666df9bca3ed494ea2674bfa1e0313e9
Nuclear RAGE interacts with the DNA replication helicase complex subunits Mcm2/7. (A) Schematic representation of the steps used for identifying the interacting partners of RAGE in the nucleus from HEK293T cells left untreated (UT) or treated with the replication stress agent hydroxyurea (HU; 2 mM for 4 h). (B) Silver-stained gel showing the proteins retrieved from the mCherry trap beads by using the lysate of untreated (–) or hydroxyurea treated (+; HU 2 mM for 4 h) HEK293T cells ectopically expressing either mCherry-RAGE or mCherry alone. Mcm2 and Mcm7 were identified by LC-MS/MS, marked by a red and black arrow, respectively. The label M indicates a protein marker. (C) Co-immunoprecipitation blots showing the interaction of RAGE with the replication helicase complex (as represented by Mcm2/Mcm7). Co-immunoprecipitation was performed from the lysate of control (–) or HU (+ represents 200μM, ++ represents 2 mM; for 4 h) treated HEK293T cells over-expressing either mCherry-hRAGE or mCherry alone. (D) Representative immunoblots showing the replication stress-induced RAGE–Mcm2 interaction. Co-immunoprecipitation was performed using mCherry trap beads from the lysate of control (–) or HU (+) (2 mM; 4 h) treated HEK293T cells over-expressing either mCherry-hRAGE and EGFP-Mcm2, or mCherry-hRAGE and alone GFP. (E) Co-immunoprecipitation immunoblots showing the RAGE–Mcm2 interaction from the lysate of control (–) or HU (+) (2 mM; 4 h) treated HEK293T cells ectopically expressing mCherry-hRAGE with anti-Mcm2 or IgG control. pCHK1 S345 served as the control to verify the activation of replication stress signaling. (F) HEK293T cells ectopically expressing either mCherry vector (lane 1), full-length mCherry-RAGE (lane 2) or mCherry-RAGE deletion constructs (lanes 3, 4 and 5) were treated with HU (2 mM, 4 h). Here FL represents full-length, ΔNTD/ΔCTD represents the deletion of the N- or C-terminal domain of RAGE, and MNBR represents the Mcm2 non-binding mutant of RAGE.
PMC10018352
gkad085fig1.jpg
0.514499
8594f1652070418197a2ac30654c7516
Role of RAGE in faithful progression of the replication fork and its stability under basal conditions or replication stress. The scheme shows the cellular response to the replication stress in the presence or absence of RAGE. The perturbed homeostasis and failure to resolve the replication errors may play a role in various diseases, such as cancer and diabetic complications.
PMC10018352
gkad085fig10.jpg
0.446161
433cfb48eeaf47b095aa71307c198491
Replication stress promotes the recruitment of RAGE into the miniature-chromosome complex. (A) The quantitative analysis showing the Mander's colocalization coefficient for the RAGE/Mcm2 or RAGE/Mcm7 colocalization of the data presented in Supplementary Figure 3A. Data represents mean ± SEM, **P < 0.01; ***P < 0.001, n = 2; >30 cells were analyzed for each condition. (B) Representative images showing the proximity ligation assay (PLA) signal in U2OS cells. These cells were ectopically expressing Flag-RAGE. Replication stress was induced by incubation with hydroxyurea (HU; 2 mM, 4 h), showing the replication stress-induced association of Mcm2 with RAGE, as marked by the PLA signal. The zoom window represents 3× magnification highlighted in dotted white boxes (scale 20 μm). (C) Schematic representation of the fluorescence-two-hybrid (F2H) system used for validating the stress-induced RAGE–Mcm2 interaction in U2OS cells (stably integrated with LacO array) by using LacI-mCherry-alone or LacI-mCherry-RAGE (bait) and EGFP–Mcm2 (prey). (D) Representative images of U2OS-LacO cells expressing the F2H system show that immobilization of RAGE onto the chromatin (bait) enriches Mcm2 (prey) onto the chromatin. The zoom window represents 2× magnification of the indicated areas, highlighted in dotted white boxes (scale 10 μm). (E) The quantitative analysis of the F2H data is shown in (D). Data represents mean ± SEM, **P < 0.01; >50 cells were analyzed for each condition. (F) Representative images showing the proximity ligation assay (PLA) signal in U2OS cells. These cells were ectopically expressing Flag-RAGE. Stress was induced by incubating them with glyoxal (0.5 mM, 4 h), showing the association of Mcm2 with RAGE, marked by the PLA signal. The zoom window represents 3× magnification of the indicated areas, highlighted in dotted white boxes (scale 20 μm).
PMC10018352
gkad085fig2.jpg
0.403106
fc2e4c1290b94cd891787aea1d0f7329
Absence of RAGE sensitizes the faithful progression of the replication fork under stress situation. (A) Representative images of the colonies from WT or RAGE−/− Hela cells either left untreated (UT) or treated with HU (indicated concentrations for 4 h), showing the percentage survival potential of these cells to recover from the HU-mediated replication stress. Colonies were stained after 10 days of recovery. (B) Quantitative analysis of the percentage survival potential of the indicated cells as described in (A). Two different Hela RAGE−/− clones were simultaneously studied. Data represented mean survival, ***P < 0.001, and ****P < 0.0001; n = 6. (C) Schematic preview and the representative images showing the pattern of symmetric and asymmetric sister forks observed in DNA fiber assay in WT or RAGE−/− Hela cells under basal conditions; scale bar, 10 μM. (D) Quantitative fiber spreading assay data showing the ratio of sister forks (shorter/longer) >50 tracts were scored for each data set, ****P < 0.0001; n = 3). The red bar indicates the median. Mann–Whitney nonparametric test was used to compare replication tract lengths and asymmetry from DNA fiber assay. (E) Schematic representation showing the plan of a modified DNA fiber assay. Here, the replication stress was induced between the pulses of CldU or IdU. (F) Representative images showing the pattern of stalled and restarted replication tracts observed in DNA fiber assay in WT or RAGE−/− Hela cells under HU treatment, scale bar, 10 μM. (G) Quantitative analysis of the fiber spreading data showing the percentage of stalled (CldU only) forks for the indicated conditions. >200 tracts were scored for each data set; *P < 0.05; **P < 0.01; n = 4.
PMC10018352
gkad085fig3.jpg
0.415073
a80cd57e2cdd4c85846501f5f4df76b2
Under replication stress situations, RAGE prevents the expression of 53BP1 in OPT domain and the appearance of micronuclei. (A) Representative immunofluorescence images showing a comparative analysis of HU stress repair potential, as evidenced by single large 53BP1 foci, as marked by yellow arrowheads, in WT or RAGE−/− Hela cells at the indicated time points (4 hours:4 h, or 16 hours:16 h). The zoom window (in dotted white boxes) represents the respective zoomed areas (scale 25 μm). (B) Quantitative analysis of the post-replication stress repair potential of the WT or RAGE−/− cells described in (A). Data represented mean ± SD, *P < 0.05; ****P < 0.0001; n = 2; >500 cells were counted for each case. (C) Representative immunofluorescence images showing the micronuclei frequency in WT or RAGE−/− Hela cells after 24 h post-replication stress (HU, 2 mM for 4 h) chase. The white arrowhead points to micronuclei (scale 25 μm). (D) Quantitative analysis of the post-replication stress repair potential of WT or RAGE−/− cells shown in (C). Data represented as mean ± SD, ****P < 0.0001; n = 3; >250 cells were analyzed for each condition.
PMC10018352
gkad085fig4.jpg
0.416765
35a406e76258498a9525748be1b3b161
RAGE stabilizes the levels of Mcm2 under replication stress. (A) Schematic presentation of the steps involved in Cycloheximidine (CHX) pulse-chase experiment in Hela RAGE−/− cells transfected with a plasmid expressing mCherry alone or the mCherry-RAGE. The samples were collected at 0, 1, 3 and 6 h. (B) Representative immunoblots showing the levels of endogenous Mcm2 in Hela RAGE−/− cells expressing mCherry alone or mCherry-RAGE treated with CHX (0.25 mM) and HU (2 mM for 4 h). The samples were collected at the indicated time points. (C) Representative immunoblots showing the levels of Mcm2 in HeLa RAGE−/− cells expressing GFP-Mcm2 in combination with mCherry alone or the mCherry-RAGE treated with CHX (0.25 mM) and HU (2 mM for 4 h) and the samples were collected at the indicated time points.
PMC10018352
gkad085fig5.jpg
0.440078
556512fc99f04153bad3159e7f489e78
Absence of RAGE affects the integrity of ciliated zones and modulates the buildup of ECM. (A) Representative images of renal sections obtained age-matched controls versus non-diabetic (3 months old) or 7-week diabetic (STZ) WT and RAGE−/− mice, stained for ciliated zone marker α-acetylated-tubulin (in red). The areas marked in dotted boxes are shown in zoomed window (scale 50μm). DAPI (in blue) represents the nucleus. (B) Quantitative analysis representing the mean fluorescence intensity of α-acetylated tubulin in WT or RAGE−/− non-diabetic (–) or diabetic (+) renal section data presented in (A). Data represent mean ± SD, **P < 0.01; ***P < 0.001, n = 5. (C) Representative images of renal sections from age-matched controls vs non-diabetic (3 months) or 7-week diabetic (STZ) WT and RAGE−/− mice showing the accumulation of ECM components, as stained by Masson's Trichrome, the ECM areas are recognized by its bluish staining (scale 50 μm). (D) Quantitative analysis representing the mean area showing accumulation of ECM components in renal sections of non-diabetic or diabetic WT or RAGE−/− mice. Data derived from (C). Data represent mean ± SD, ****P < 0.0001; n = 5.
PMC10018352
gkad085fig6.jpg
0.454519
734bcaeaeaeb49a395e412127542dc1f
RAGE and Mcm2 participate in the same pathway to counter the replication stress. (A) Quantitative analysis showing the percentage survival potential of untreated (UT; black bars) and hydroxyurea treated (HU; 2 mM, 4 h; grey bars), Mcm2 depleted WT or RAGE−/− HeLa cells as described in Supplementary Figure 20B. Data represent mean ± SD, ****P < 0.0001; n = 6. (B) Representative images showing the pattern of 53BP1 (in red) stained after 16 h of post-recovery of HU (2 mM, 4 h) treated and Mcm2 depleted WT or RAGE−/− Hela cells. DAPI (in blue) represents the nucleus, and the zoom window (in dotted white boxes) represents the respective 3x zoomed areas (scale 25 μm). (C) Quantitative analysis showing the mean percentage of 53BP1 marked OPT domains from the cells described in (B). Data represent mean ± SD, *P < 0.05, ****P < 0.0001, ns represents non-significant; n = 3. In this figure, ns represents the comparison between single depletion/deletion, whereas ‘ns’ indicates double depletion/deletion. (D) The quantitative analysis showing the frequency of micronuclei from HU-treated cells. Data represent mean ± SD, *P < 0.05, **P < 0.01; ***P < 0.001; ****P < 0.0001; n = 3.
PMC10018352
gkad085fig7.jpg