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
0.53812
10a4c9f9a46d42adbed33c3eef64e618
Flowchart.
PMC9072350
41598_2022_11346_Fig1_HTML.jpg
0.400273
a3ca783117974fdd839cdf9aa569fcec
Wexner score and mode of delivery.
PMC9072350
41598_2022_11346_Fig2_HTML.jpg
0.495152
731dd2aad6a446c5adbf193d23919bc8
Temperature–time profile with three measurement stages: (i) 1st stage: sample mass determination (red segment), (ii) 2nd stage: sample melting and fast-quenching (green segment), and (iii) 3rd stage: re-heating of supercooled sample (blue segment). After cooling step #4 in 1st stage, the sample can be coated with silicon oil. In the heating step #5, the scanning rate, β, varied from 2000 K s−1 to 20 000 K s−1 [reprinted from ref. 24 with modifications].
PMC9073158
c9ra05730g-f1.jpg
0.424818
3eb5a73424cf4d9f9868d0a1791995fd
Extrapolated onset temperature of the melting peak of Gly–Gly, Gly–Ala, Ala–Gly, Ala–Ala and cyclo(Ala–Gly), as function of heating rate. The melting temperature at zero heating rate for Gly–Gly, Gly–Ala, Ala–Gly, Ala–Ala and cyclo(Ala–Gly) is TSL0Gly–Gly = (593 ± 7) K, TSL0Gly–Ala = (551 ± 7) K, TSL0Ala–Gly = (611 ± 7) K, TSL0Ala–Ala = (606 ± 7) K and TSL0cyclo(Ala–Gly) = (526 ± 7) K, respectively. The scanning rates used were 2000 K s−1 (circles), 4000 K s−1 (up-triangles), 5000 K s−1 (hexagonals), 6000 K s−1 (down-triangles), 8000 K s−1 (diamonds), 10 000 K s−1 (stars), 15 000 K s−1 (left-triangles) and 20 000 K s−1 (right-triangles). Solid symbols represent sample measurement without silicon oil, while empty symbols for sample measurement with silicon oil.
PMC9073158
c9ra05730g-f2.jpg
0.465893
9016e9ae2e424979acac66842dd3ad39
Enthalpy, ΔHSL0i, of Gly–Gly, Gly–Ala, Ala–Gly, Ala–Ala and cyclo(Ala–Gly) in respect to initial sample mass, m0, regardless of the scanning rates. The symbols used were as described in Fig. 2. The dashed line was linear fit through zero origin, where the slope denoted as ΔhSL0i. The scanning rates used were 2000 K s−1 (circles), 4000 K s−1 (up-triangles), 5000 K s−1 (hexagonals), 6000 K s−1 (down-triangles), 8000 K s−1 (diamonds), 10 000 K s−1 (stars). The melting enthalpy for Gly–Gly, Gly–Ala, Ala–Gly, Ala–Ala and cyclo(Ala–Gly) is ΔhSL0Gly–Gly = (40 ± 6) kJ mol−1, ΔhSL0Gly–Ala = (41 ± 5) kJ mol−1, ΔhSL0Ala–Gly = (52 ± 7) kJ mol−1, ΔhSL0Ala–Ala = (45 ± 7) kJ mol−1 and ΔhSL0cyclo(Ala–Gly) = (24 ± 4) kJ mol−1, respectively. The values are already given in the figures, same for Fig. 2.
PMC9073158
c9ra05730g-f3.jpg
0.413909
9b3db577b3df43fb842f999390755c7a
Specific heat capacity for Gly–Gly, Gly–Ala, Ala–Gly and Ala–Ala. The heat capacity of solid, cSp0i, of the dipeptides was measured with conventional DSC (dotted lines). The solid line denotes the glass transition step of ultra-fast quenched melted dipeptides without silicon oil. Both cSp0i and cLp0i (dashed lines) were linearly fitted to extrapolate to TSL0i. The heat capacity difference between liquid and solid phase were determined at glass transition temperature, ΔcSLp0i (TG0i) and at melting temperature, ΔcSLp0i (TSL0i). Gly–Gly: ΔcSLp0Gly–Gly (TG0Gly–Gly) = (84 ± 6) J mol−1 K−1 and ΔcSLp (TSL0Gly–Gly) = (51 ± 6) J mol−1 K−1; Gly–Ala: ΔcSLp0Gly–Ala (TG0Gly–Ala) = (91 ± 6) J mol−1 K−1 and ΔcSLp0Gly–Ala(TSL0Gly–Ala) = (55 ± 6) J mol−1 K−1; Ala–Gly: ΔcSLp0Ala–Gly (TG0Ala–Gly) = (82 ± 3) J mol−1 K−1 and ΔcSLp (TSL0Ala–Gly) = (57 ± 3) J mol−1 K−1; Ala–Ala: ΔcSLp0Ala–Ala (TG0Ala–Ala) = (84 ± 18) J mol−1 K−1 and ΔcSLp0Ala–Ala (TSL0Ala–Ala) = (62 ± 18) J mol−1 K−1. The solid squares depict specific heat capacity of solid Gly–Gly from literature.60 In order to avoid crystallization on cooling, higher scanning rates are required and might be able to achieved with ultra-fast scanning nanocalorimetry.61,62
PMC9073158
c9ra05730g-f4.jpg
0.424378
9219d77d18c54ce7be8e11acc89d7096
Experimental dipeptide solubility at pH = 7 in water as molality vs. temperature. Full symbols present data measured by photometric method; empty symbols by gravimetric method. Gly–Ala (squares + “x”-filled square38); Ala–Ala (up-triangles); Gly–Gly (circles);11 Ala–Gly (down-triangles) and cyclo(Ala–Gly) (diamonds). The experimentally determined values are given in Tables S5 and S6 (pH 7) and Tables S7 and S8 (pH at saturated solutions) in the ESI.†
PMC9073158
c9ra05730g-f5.jpg
0.505546
833e8b7543c84a76b3f358c9f7f527fe
Solubility in water. Bars: mean values of the photometric and gravimetric determined dipeptide solubility data at T = 298.15 K and pH = 7. Lines represent the corresponding amino-acid solubility at T = 298.15 K and pH = 7: solid line: Gly24 and dashed line: Ala.24
PMC9073158
c9ra05730g-f6.jpg
0.437879
77514dd80dbc43a49287f9528363d896
Influence of the difference of the heat capacities on the solubility behaviour for Gly–Gly (circles11) in water as molality vs. temperature. Lines represent PC-SAFT predictions with the parameters from Table 3 and FSC-measured melting properties from this work (Table 4). Dotted line: eqn (1) with ΔcSLp0i = 0, dotted-dashed line: eqn (1) with ΔcSLp0i = const. = 51 J mol−1 K−1, solid line: eqn (1) with eqn (2).
PMC9073158
c9ra05730g-f7.jpg
0.471477
6b79da6c897042f0a963d5efffe840f1
Influence of the activity coefficient on the solubility in molality in water expressed as difference between ideal and experimental solubility. Black (Ideal): calculated solubility at T = 298.15 K from eqn (1) based on the FSC-measured melting properties in experimental uncertainty assuming an ideal mixture γ = 1. Shaded (Exp.): mean value of the photometric and gravimetric determined solubility data at T = 298.15 K and pH = 7. Grey (PC-SAFT): calculated solubility at T = 298.15 K from eqn (1) based on the PC-SAFT used melting properties from Table 4 including the activity coefficient.
PMC9073158
c9ra05730g-f8.jpg
0.447584
ba96357a46864990a2c5bbe58a82fde8
Dipeptides solubility at pH = 7 in water as molality vs. temperature. Symbols represent experimental data. Solid symbols present measurements using photometric method; open symbols present measurements using gravimetric method. Gly–Ala (squares + "x" filled square38); Ala–Ala (up-triangles); Gly–Gly (circles11); Ala–Gly (down-triangles); cyclo(Ala–Gly) (diamonds). Lines represent PC-SAFT predictions with the parameters from Table 3 and FSC-measured melting properties from this work. Gly–Ala (solid line), Ala–Ala (dashed line), Gly–Gly (dashed-dotted line), Ala–Gly (dashed-double dotted line) and cyclo(Ala–Gly) (dotted line). The experimental determined values are given in Tables S5 and S6 (pH = 7) and Tables S7 and S8† (pH at saturated solutions).
PMC9073158
c9ra05730g-f9.jpg
0.442068
e71609a07ad34acfbcb53ff97be8a83a
Graphical example of Hill–Robertson effects. Hitchhiking effects associated with linkage to a deleterious mutation, known as background selection, may act to reduce variation, as purifying selection acting on the deleterious mutation (shown in red) may result in the elimination of linked variants. Similarly, hitchhiking effects associated with linkage to a beneficial mutation, known as a selective sweep, may also act to reduce variation, as positive selection acting on the beneficial mutation (shown in blue) may result in the fixation of linked variants.
PMC9073693
jkac055f1.jpg
0.440706
7dd7edd2d50748c78e8f9140aedce849
Comparison of within-host variation between M. canettii and MTBC. Mycobacterium canettii has both a greatly elevated mean and variance in genome-wide segregating sites relative to the MTBC.
PMC9073693
jkac055f2.jpg
0.472174
f81a9942d06f496cbee8cc5af068310c
Distribution of likelihoods across different values of ρ for empirical M. canettii. Distribution of likelihood for different population-scaled recombination rates, ρ, ranging from 0 to 100 obtained using LDhat (see Materials and Methods for details). The blue line indicates the maximum likelihood at ρ = 15.
PMC9073693
jkac055f3.jpg
0.447296
ba489e12750544278f4c0d026688e84a
Performance of LDhat under a variety of model violations. a) the per-site recombination rate (r) inferred from ρ estimated by LDhat deviates strongly from the true value of r = 7.2 × 10−11 (see solid black horizontal line) in most models. Indeed, the correct estimate is obtained only when using full population data (rather than consensus data) under either the standard Wright–Fisher (Base) model or a neutral bottleneck model (Base + Bn). Each point corresponds to a simulated replicate, with replicates binned according to their model along the x-axis. The point color corresponds to the ratio between the LDhat-inferred r and the true r (log-scaled to improve visualization). b) Similar to panel a, the relationship between ρ and θ deviates from the slope of r/μ (solid black line) for many of the simulated model violations (colors represent different models). Both the standard Wright–Fisher (Base) and the neutral bottleneck (Base + Bn) models largely conform to expectation when the full population data are used. However, the use of consensus sequences as well as the presence of either natural selection and/or progeny skew, can cause extreme deviations from the true value.
PMC9073693
jkac055f4.jpg
0.411413
5d86b5e17e474d769c840aecb02a6c95
Segregating sites in consensus sequences and source populations, from simulated replicates. The number of segregating sites observed within each consensus, full population, and subpopulation dataset—for 10 simulation replicates of the ‘Base’ model with the addition of a bottleneck, progeny skew, and presence of non-neutral mutations (Base + Bn + Ψ + DFE). The full populations are represented by 500 genomes, the subpopulations are represented by 17 subpopulations consisting of 25 individual genomes each, and the consensus sequences are called for each subpopulation from the 25 constituent individual genomes. The points are binned by simulation replicates.
PMC9073693
jkac055f5.jpg
0.500209
c2255122e30a45fd96c213c2d1c94ef6
PRISMA flow chart of the scoping review. Summary of evidence search and selection
PMC9074253
12935_2022_2603_Fig1_HTML.jpg
0.460159
5c1059800533442fa93690a704612558
Scamper diagram showing Mcl-1 expression in oral cavity cancers and association with various clinicopathological features
PMC9074253
12935_2022_2603_Fig2_HTML.jpg
0.493743
69d10752ca80458591ff235c76e87bdb
STRING protein–protein interaction (PPI) analyses. PPI network connectivity for proteins identified following the review. Nodes represent the proteins required for interaction. Edges represent the associations between the proteins. The STRING web resource (http://www.stringdb.org) was used in the prediction of the PPI (Protein–Protein Interaction) network whereby an interaction score of > 0.900 denoted a significant interactive relationship
PMC9074253
12935_2022_2603_Fig3_HTML.jpg
0.431577
357fd608ee3247159aec77a94a746a70
Tree modified from Figure 5 in Andrade et al. (2015) including the species covered in this study. Note that some of the internal relationships of the annelids are not yet well resolved, so this figure only gives an overview of the diversity covered within the group.
PMC9075518
fphys-13-817272-g001.jpg
0.508898
f8559cd915494f63902acf35be6293f9
Flow chart of identification and analysis of sHsp.
PMC9075518
fphys-13-817272-g002.jpg
0.402909
6fc466f3a44e4edbb7c9bbccb48482af
Representation of the structural topology of monomeric sHsps. ACD is the region delimited from the β2-strand to the β9-strand. The β1-strand is localized in the N-terminal region. The β10-strand corresponds to a conserved motif in the C-terminal region. The fragment starting after the ACD and including the β10-strand has been named the C-terminal anchoring module (CAM). Residues that follow the CAM were defined as the C-terminal tail. The β6-strand and the β10-strand are not universally present in all sHsp (Poulain et al., 2010). The black arrows represent the β-strands and the red line, the link between two β-strands.
PMC9075518
fphys-13-817272-g003.jpg
0.438801
3fd0cad1d13a439fbeb2ecfe70b1bc0f
ML tree generated using the multiple sequence alignment of the ACD data set (A) Left, unrooted ML phylogenetic tree and clustering branching pattern of monomeric sHsps (shadowed clades) (B) Right, circular ML phylogenetic tree showing the location of the outgroups’ ACDs (shadowed labels), the colored connection lines between the ACDs belonging to the same dimeric protein, and the colored stripes (green, red and blue) marking the three big clades: Cluster A, B and C, respectively. Branch lengths are not displayed. More details can be found in File S2, including the results of the Bayesian phylogenetic study, the dendrograms of the analyzed clusters, and the node support information.
PMC9075518
fphys-13-817272-g004.jpg
0.406629
cccc130984fd4967943e045008dbff03
Clusters A1 and A2. Multiple sequence alignment visualized on a dendrogram obtained from the Bayesian tree, using the Zappo coloring scheme. The consensus sequence (at 50% conservation) and residue conservation were calculated by iTOL. On the left, the unrooted radial ML cladograms, in which each cluster is highlighted. The posterior probability values are included as percentages in the nodes.
PMC9075518
fphys-13-817272-g005.jpg
0.362226
b669af654cbf4d6b901c9962d3b755d6
Clusters B1, B2, and B3. Multiple sequence alignment visualized on a dendrogram obtained from the Bayesian tree, using the Zappo coloring scheme. The consensus sequence (at 50% conservation) and residue conservation were calculated by iTOL. On the left, the unrooted radial ML cladograms, in which each cluster is highlighted. The posterior probability values are included as percentages in the nodes.
PMC9075518
fphys-13-817272-g006.jpg
0.486062
b52e13c598974ee7b12f14db25ce4a45
Logo presentation using the WebLogo3 program (Crooks et al., 2004) for Clusters A and B. The height of each letter is proportional to its frequency. Amino acids are colored according to their chemical properties: acidic (D, E) in red, basic (K, R, H) in blue, polar (G, S, T, Y, C) in green, neutral (N, Q) in purple, and hydrophobic (A, V, L, I, p, W, F, M) in black. Logos for Clusters B2 and B3 are calculated without including the sequences from the outgroups (Supplementary File S2, Supplementary Figure S17-S19 include logos for Clusters B2 and B3 calculated including the outgroups, but the differences are not significant.).
PMC9075518
fphys-13-817272-g007.jpg
0.5317
cb79b38aa3e9423fab0df496ba071a3d
THP chelator (R = CH3) and its bifunctional chelator derivatives.
PMC9075519
c9ra07723e-f1.jpg
0.44185
901d277a140a4934b67cfab968e6574c
Concentrations of selected trace metals in 68Ga generator eluate measured by ICP-MS. Generator eluate (5 mL in 0.1 M HCl) was obtained from a single generator either 2 hours after a previous generator elution (sample set A, red circles, n = 9), or 1 day after a previous elution (sample set B, purple squares, n = 4). “Blank” eluent (0.1 M HCl solution from ABX Gmb) was also measured (yellow triangles, n = 10). These samples were collected within a period of 20 days, from a generator eluting 330–480 MBq. Error bars represent standard deviation of concentrations in the samples.
PMC9075519
c9ra07723e-f2.jpg
0.383321
8cdc1354ce3b429983b20c5305980515
Concentrations of Al, Ti, Fe and natGa in 68Ga eluate as a function of generator age. Two sets of eluate samples were collected from the same 68Ga E&Z generator, six months apart from each other. “Blank” eluent (0.1 M HCl solution from ABX Gmb) used in these two sets of samples was also measured. During this period, the generator was eluted 130 times. All eluates were sampled with a pre-elution window of 2 h. Error bars represent standard deviation of concentrations in the samples.
PMC9075519
c9ra07723e-f3.jpg
0.41308
2b677e0d29bb415ead650aa019f61558
Mean RCYs of [68Ga(THP)] in the presence of increasing concentrations of metal ions ((A): Al3+, Ti4+, Cr3+, Fe3+ and (B): Ni2+, Zn2+, natGa3+, Pb2+). The vertical dashed line indicates equimolar concentrations of chelator and metal ion (5 μM), and the horizontal dashed guide indicates 97% (n = 7) RCY, which is achieved in the absence of added metal ions. For most values, error bars representing standard deviation are smaller than the symbols. For a full list of mean RCY (n = 3) and standard deviation values see Table SI-4.†
PMC9075519
c9ra07723e-f4.jpg
0.475747
62932e1bffc14eeabf23f8d152c9d6e7
Mean RCYs of [68Ga(THP)] in the presence of either increasing concentrations of Fe3+ (green triangles) or ascorbate (166 mM) and increasing concentrations of Fe3+ (n = 5). The vertical dashed line indicates equimolar concentrations of chelator and metal ion (5 μM), and the horizontal dashed guide indicates 96 ± 0.4% (n = 6) RCY in the absence of spiked Fe3+. For most values, error bars representing standard deviation are smaller than the symbols.
PMC9075519
c9ra07723e-f5.jpg
0.559379
309a525cfaa5427ca549c5e20146010a
 Regression lines, with 95% confidence intervals, for maternal (R 2  = 0.5813) and umbilical cord blood (R 2  = 0.7113) IgG seropositivity relative to the infection-to-birth interval (weeks). Blue circles indicate percentages (Y axis) of mothers, while green dots represent percentages of umbilical cord blood testing positive for anti-SARS-CoV-2 IgG for the indicated weeks between infection and birth (X axis). The regression lines differed significantly (comparison of fits: p < 0.0001, F  = 8.958; dfn  = 4, dfd  = 340). Note the delayed response in seroconversion by one week between mothers and their offspring early after infection.
PMC9076216
10-1055-a-1768-0415-igf01.jpg
0.437445
7d79b7f619dc48318f749419f6aa303c
 Scatter plot and regression line for logarithmic maternal and logarithmic umbilical cord IgG concentrations. Each dot represents a mother–child pair, both of whom tested positive for anti-SARS-CoV-2 IgG antibodies (n = 68, ρ = 0.8042; p < 0.0001).
PMC9076216
10-1055-a-1768-0415-igf02.jpg
0.509014
f2b58ed0490e4d3fab9ae0bba4dcfbf6
 Third-order cubic polynomial regression line for maternofetal anti-SARS-CoV-2 IgG antibody transfer ratios (c [umbilical cord blood IgG]/c [maternal blood IgG]) relative to the infection-to-birth interval. Note that the ratios increase to more than 1 after six weeks from infection (R 2  = 0.2440).
PMC9076216
10-1055-a-1768-0415-igf03.jpg
0.467892
0e387dd096a0483482a876986d492624
1H NMR spectra and their assignments of PLLA2–PBLG (B) and PLLA2–PLGA (A) in DMSO-d6 (x and y are solvent peaks).
PMC9076388
c9ra08703f-f1.jpg
0.40938
d53ccb8457eb40d5a6df5181bf103aa7
TEM micrographs of the lyophilized vesicles from PLLA2(30)–PLGA55.
PMC9076388
c9ra08703f-f10.jpg
0.521757
c9efeaa264c84a58bf885a2827da4335
IR spectra of (A) PLLA2-b-PBLG, (B) PLLA2-b-PLGA.
PMC9076388
c9ra08703f-f2.jpg
0.581348
1a9b5d60997e4a2c80b133ea8923068a
The GPC chromatographs of PLLA2–NH–Z (b), PLLA2(30)–PBLG10 (c) and PLLA2(30)–PBLG55 (a).
PMC9076388
c9ra08703f-f3.jpg
0.48087
9d1f377fc8bc4ba088e9c0291aa96b84
(A) Plot of I333/I331 of pyrene vs. log C of PLLA2(30)–PLGA55 in deionized water. (B) Plot of I335/I332 of pyrene vs. log C of PLLA2(30)–PLGA10 in deionized water.
PMC9076388
c9ra08703f-f4.jpg
0.428927
f5c5aaebdd9a4e93a7f259ed01aced9d
(A) TEM micrographs of the vesicles from PLLA2(30)–PLGA55. (B) TEM micrographs of the vesicles from PLLA2(30)–PLGA10. (C) ESEM micrographs of vesicles from PLLA2(30)–PLGA55, the inset is magnification. (D) Schematic representation of the vesicle.
PMC9076388
c9ra08703f-f5.jpg
0.395256
8ec2de255a0d41c5a2639c7b512bdb45
DLS graphs of the micelle size distribution of (A) PLLA2(30)–PLGA55 and (B) PLLA2(30)–PLGA10.
PMC9076388
c9ra08703f-f6.jpg
0.440735
faeae39bfc7747faa8578ad2c7b57abf
1H NMR spectra and their assignments of PLLA2(30)–PLGA10 in D2O.
PMC9076388
c9ra08703f-f7.jpg
0.527612
e66e8ca501f34774ad3883e16e208f76
Plot of Rh of PLLA2(30)–PLGA55 vesicles vs. solution pH value (the concentration of the copolymer is 0.4 mg ml−1).
PMC9076388
c9ra08703f-f8.jpg
0.467491
e59d389c8bd845f581d048841049d279
Plot of Rh of PLLA2(30)–PLGA55 vesicle vs. NaCl concentration at pH 5.5 (the concentration of the copolymer is 0.4 mg ml−1).
PMC9076388
c9ra08703f-f9.jpg
0.433006
5f2a7853841740d580757a3b476673f0
The aldehyde content of NCDs at pHs 3 and 7 versus oxidation times.
PMC9076516
c9ra05240b-f1.jpg
0.464307
6e718e1b38674d87b096233d1a430234
FT-IR spectra of MC, HANCD, and immobilized urease as HANCD@urease.
PMC9076516
c9ra05240b-f2.jpg
0.51657
0a33326fa39c4c179a1bcbaf0a1e2f42
X-ray diffraction patterns of HANCD and urease immobilized on HANCD.
PMC9076516
c9ra05240b-f3.jpg
0.396271
5b4a58bc2dca4e268dc1a6c6d5882cfc
FESEM of MC (a), NC (b), HANCD (c), and immobilized urease on HANCD (d).
PMC9076516
c9ra05240b-f4.jpg
0.504123
0072a852659e41e5a2844073a6a9cca5
Thermal gravimetric analysis of HANCD and HANCD@urease.
PMC9076516
c9ra05240b-f5.jpg
0.466162
f270b2bd00f24f0d8989736b8df389d4
The relative residual activities of the free and immobilized urease at adjusted pHs at 35 °C (a) and at temperatures from 20–80 °C at pH = 7 (b); the aldehyde content of HANCD@urease at various temperatures and pHs (c); change in the absolute activities for free and HANCD anchored urease at various pHs (d) and temperatures (e). Each experimental point represents the mean of five determinations with 2.8% SD.
PMC9076516
c9ra05240b-f6.jpg
0.482001
1e0aa068b88641f4a7e530b2ef624ef4
Comparative residual activity of the free- and immobilized-urease versus time.
PMC9076516
c9ra05240b-f7.jpg
0.444327
c903e81b10394ca4a54b7d47c92da94a
Operating reusability of the HANCD@urease evaluation of serum urea by HANCD@urease.
PMC9076516
c9ra05240b-f8.jpg
0.424127
04d0fbf0aefc4d6b80f464aedd7d5445
CircPVT1 and NEK7 were up-regulated while miR-181a-5p was down-regulated in NSCLC tissues and cells. (A–D) QRT-PCR was implemented to measure the expression of circPVT1 (A and B) and miR-181a-5p (C and D) in NSCLC tissues and cells. (E–H) The mRNA and protein levels of NEK7 in NSCLC tissues (E and F) and cells (G and H) were detected through qRT-PCR and western blot. (I–K) The linear relation among circPVT1, miR-181a-5p and NEK7 was analyzed by Spearman's correlation coefficient. *P < 0.05.
PMC9076563
c9ra08872e-f1.jpg
0.38639
6bbc8f4b9c184ec4a9e15725180faa9b
Knockdown of circPVT1 suppressed the chemoresistance of cisplatin and metastasis in NSCLC cells. (A) The interference efficiency of si-circPVT1 was assessed by qRT-PCR in NSCLC cells. (B and C) Cell viability was examined using an MTT assay after transfection with si-circPVT1 or si-NC (B) and IC50 was calculated (C). (D and E) Cell apoptosis was evaluated through flow cytometry (D) and the detection of apoptosis-associated proteins by western blot (E) after treatment with 2 μg mL−1 cisplatin and transfection. (F and G) The ability of metastasis was determined via a transwell invasion assay (F) and the levels of EMT-related proteins by western blot (G). *P < 0.05.
PMC9076563
c9ra08872e-f2.jpg
0.416666
8fbe745751624e60acdabb1630d5cf6a
CircPVT1 induced targeted regulation of the level of miR-181a-5p. (A) The bioinformatics analysis between circPVT1 and miR-181a-5p was conducted via Starbase2.0. (B and C) The relationship between circPVT1 and miR-181a-5p in NSCLC cells was verified by a dual-luciferase reporter assay (B) and RIP assay (C). (D) QRT-PCR was applied for assaying the effects of circPVT1 inhibition or overexpression on the miR-181a-5p expression in A549 and H1299 cells with si-NC and pcDNA-NC as the negative controls. *P < 0.05.
PMC9076563
c9ra08872e-f3.jpg
0.49208
716cec209d424aa1910fcb050916c4dc
Down-regulation of circPVT1 ameliorated the effects of miR-181a-5p inhibitor on NSCLC cells. (A) The miR-181a-5p expression was assayed by qRT-PCR in A549 and H1299 cells transfected with miR-NC inhibitor, miR-181a-5p inhibitor, miR-181a-5p inhibitor + si-NC or miR-181a-5p inhibitor + si-circPVT1. (B and C) MTT was adopted for examining cell viability and determining the value of IC50 of cisplatin after transfection with miR-181a-5p inhibitor, miR-181a-5p inhibitor + si-circPVT1 or relative controls. (D and E) Flow cytometry and western blot were used for the determination of the apoptosis rate and related protein expression in A549 and H1299 cells treated with 2 μg mL−1 cisplatin and transfected with the above groups, respectively. (F) A transwell assay was applied to detect the ability of invasion. (G and H) The levels of EMT-associated markers were measured using western blot. *P < 0.05.
PMC9076563
c9ra08872e-f4.jpg
0.470562
c1493e9474164823a04e094fcc2fe170
Inhibition of circPVT1 aggravated the si-NEK7-inhibited metastasis via reducing NEK7 by promoting miR-181a-5p. (A) The prediction of the targets of miR-181a-5p was executed using Starbase2.0. (B and C) A dual-luciferase reporter assay and RIP assay were used for validating the target relation between miR-181a-5p and NEK7 in A549 and H1299 cells. (D and E) The mRNA and protein levels of NEK7 were measured by qRT-PCR and western blot after transfection with si-NEK7, si-circPVT1, miR-181a-5p inhibitor, si-circPVT1 + miR-181a-5p inhibitor or matched controls. The expression levels of circPVT1 and miR-181a-5p were determined in A549 and H1299 cells transfected with si-NEK7 or si-NC. (H–J) Cell metastasis was assessed by transwell invasion assay and the assaying of EMT-related markers through western blot in A549 and H1299 cells transfected with si-NEK7, si-NEK7 + si-circPVT1, si-NEK7 + miR-181a-5p inhibitor or matched controls. *P < 0.05.
PMC9076563
c9ra08872e-f5.jpg
0.471692
8852216764e2423b9864396c695df91f
Knockdown of circPVT1 relieved the 3-MA-promoted chemoresistance of cisplatin via elevating miR-181a-5p-mediated autophagy. (A) The level of miR-181a-5p was determined by qRT-PCR in A549 and H1299 cells treated with 3-MA, 3-MA + si-circPVT1, 3-MA + miR-181a-5p inhibitor or respective controls. (B) Western blot was used for assaying the protein levels of autophagy-related proteins. (C and D) Cell viability and IC50 of cisplatin were measured and analyzed by MTT. (E and F) After treatment with 2 μg mL−1 cisplatin and treatment with the above groups, cell apoptosis was assessed from the apoptosis rate by flow cytometry and apoptosis-associated protein expression by western blot. *P < 0.05.
PMC9076563
c9ra08872e-f6.jpg
0.453982
fd7abb77a7514979bd9be7004cefdc5f
The decrease of circPVT1 heightened the cisplatin sensitivity of NSCLC cells in vivo. (A and B) Tumor volume and weight were calculated in four groups of sh-NC + PBS, sh-NC + cisplatin, sh-circPVT1 + PBS and sh-circPVT1 + cisplatin. (C) The expression of circPVT1 was detected by qRT-PCR using the RNA from the tumor tissues. *P < 0.05.
PMC9076563
c9ra08872e-f7.jpg
0.499405
907ffd04924a435496f356e8fa439823
The regulatory mechanisms of circPVT1 on metastasis and chemoresistance of NSCLC. CircPVT1 interacts with miR-181a-5p to regulate NEK7 expression, affecting cell metastasis, or regulate autophagy, promoting chemoresistance.
PMC9076563
c9ra08872e-f8.jpg
0.460742
adfabe5177c445c0852725dae934bebd
Circular dichroism and cell viability analysis of DE-11. Peptide DE-11 was dissolved in 20 mM HEPES (pH = 7.4). CD spectra of DE-11 after 2 h (full line) and 24 h (broken line) of incubation (A). CD spectra of DE-11 after the addition of CaCl2 (red line) and Na2HPO4 (green line), in comparison with those of DE-11 without Ca2+ and HPO42− (black line) (B). The effect of HOK cell viability of DE-11 at various concentrations determined by the CCK-8 assay. Data are presented as the mean ± SD (C).
PMC9077281
c7ra12032j-f1.jpg
0.409486
c5a9bc37ef314c47a57b1946d259bcf9
Linear adsorption isotherms of DE-11. Protein–HA adsorption data fit the Langmuir model well. The affinity of DE-11 molecules for HA adsorption sites (K) and the maximum number of adsorption sites per gram of HA (N) were calculated. R2 is the correlation coefficient obtained for linear adsorption isotherms.
PMC9077281
c7ra12032j-f2.jpg
0.441866
9f9d528d64734c9c87085645de4c16a1
SEM images of calcium phosphate minerals in the control group (A), DE-11 group (B), and DK-6 group (C) after 24 h of incubation. Spherical particles were formed in both the DK-6 and the calcium phosphate control samples, whereas a different morphology characterized as smaller crystals with rough and spicular surfaces was observed in the DE-11 sample.
PMC9077281
c7ra12032j-f3.jpg
0.397506
704342ce23df4b1784303514f68da737
TEM images of calcium phosphate minerals and related SAED patterns after 24 h of incubation. Calcium phosphate control group: morphology (A) and SAED pattern (B). DE-11 group: morphology (C) and SAED pattern (D). DK-6 group: morphology (E) and SAED pattern (F). The mineral structure consisting of needle-like crystals was clearly observed in the presence of DE-11.
PMC9077281
c7ra12032j-f4.jpg
0.445424
c1ef21555890426086af7d40aa1c7713
SMH analysis of the enamel samples after the remineralization assay. The SMH of the NaF group, DE-11-treated group, DK-6-treated group, and HEPES group after immersion in artificial saliva for 3 and 7 days as compared to that of the original enamel samples (A). SMHR% of the NaF group, DE-11-treated group, DK-6-treated group, and HEPES group after being immersed in artificial saliva for 3 days (B) and for 7 days (C). Data are presented as the mean ± SD. The statistical analysis was performed using ANOVA followed by the Student–Newman–Keuls test (**P < 0.01, ***P < 0.001).
PMC9077281
c7ra12032j-f5.jpg
0.465029
d77dd78cef874cc28ec122358e594569
PLM images of the enamel sections before and after remineralization treatment in the presence of NaF (A), DE-11 (B), DK-6 (C) or HEPES alone (D).
PMC9077281
c7ra12032j-f6.jpg
0.457644
294dcaa6b008489c8fac230035bdb20a
TMR analysis of enamel sections before and after remineralization treatment in the presence of NaF, DE-11, DK-6 or HEPES alone. Mineral loss of enamel blocks (A), lesion depth of enamel blocks (B), and mineral content (vol% μm) vs. depth (μm) for lesions (C). Data are presented as the mean ± SD. The statistical analysis was performed using the Students paired t test. Bars labelled with different letters are significant differences (P < 0.05).
PMC9077281
c7ra12032j-f7.jpg
0.456892
13396ecb0d5f4142adde16410c543d72
Schematic of the adsorption of DE-11 on the demineralized enamel surface for remineralization restoration.
PMC9077281
c7ra12032j-f8.jpg
0.436376
d5ae7ef9960f4884b180d2c81afed09d
The experimental device and electrode positions.
PMC9078048
iovs-63-5-7-f001.jpg
0.41295
4c595e28a82f4447a0b049b32c9d4dc3
The maximal power of theta-, alpha- and beta-waves of 13 electrodes of ROI in condition 1 (20″, unperceived totally) and condition 2 (100″, perceived clearly). (a) Heat map; (b) summary data and ANOVA analysis.
PMC9078048
iovs-63-5-7-f002.jpg
0.472712
0b31f00e4f8d4f6380f0ff3678b7a540
The changes (∆) of maximal powers of theta-, alpha-, and beta-waves of the 13 electrodes in the ROI induced by depth perception. (a, b) The ∆ maximal powers of theta-, alpha-, and beta-waves in all the electrodes. (c) Grid of Tukey's multiple comparisons of ∆ maximal powers of the theta-waves. (d) Grid of Tukey's multiple comparisons of ∆ maximal powers of the alpha-waves.
PMC9078048
iovs-63-5-7-f003.jpg
0.442059
8705c5a1946b4c7ab1f8e2c6568713c9
Functional connectivity of theta and alpha-waves between any two channels among the interesting electrodes from the frontal lobe and occipital lobe. (a) The PLI values of theta-wave under condition 1 (20″, unperceived totally). (b) The PLI values of the theta-waves under condition 2 (100″, perceived clearly). (c) The FC changes of the theta-waves and the comparisons between FCs in condition 2 and those in condition 1. (d) The PLI values of the alpha-waves under condition 1. (e) The PLI values of the alpha-waves under condition 2. (f) The FC changes of the alpha-waves and the comparisons between FCs in condition 2 and those in condition 1 (* P < 0.05, # P < 0.01, + P < 0.001).
PMC9078048
iovs-63-5-7-f004.jpg
0.491501
e1a9ae49adab4c949325f401f5b673f1
Ball-and-stick structures of compounds Q7 and Q9.
PMC9078235
c7ra13397a-f1.jpg
0.470694
c40d38023d3e41868ad78c11e2c1e22f
Thermogravimetric and heat flow curves of compounds Q1–Q9 recorded at a constant heating rate. The mass loss is expressed as the fraction M/M0 of the mass at temperature T to the initial mass of the samples equilibrated at laboratory atmosphere.
PMC9078235
c7ra13397a-f2.jpg
0.409414
3febde8bf2c24fed9f31d00c0031deff
The EC50 values for the activation of thrombin activity by compounds Q2 and Q8. aThe EC50 values are expressed as the concentration required for a thrombin activity response of 50%; n = 3.
PMC9078235
c7ra13397a-f3.jpg
0.444132
19a16eadd5b34a57afd2c9092c0e34f5
The classical systems for blood coagulation including extrinsic, intrinsic, and/or the common coagulation pathways.30
PMC9078235
c7ra13397a-f4.jpg
0.473026
5fc16f63d6624fbcb669c35d8a5ac142
Compound Q2 docking into the thrombin binding pocket.
PMC9078235
c7ra13397a-f5.jpg
0.423433
977a81944f454280848aa9fc09d11668
Effects of compounds Q2 and etamsylate on capillary permeability. The results were expressed as the concentration of CNP (pg mL−1). Values are reported as the mean ± SEM of three independent experiments.
PMC9078235
c7ra13397a-f6.jpg
0.43741
7dc204242d7e4a2799ce115b8cc8564d
(a) XRD patterns of La0.65Ce0.05Sr0.3Mn1−xCuxO3 (0 ≤ x ≤ 0.15) compounds at room temperature. (b) Rietveld refinement profile for x = 0.05 performed using FULLPROF. Open circles correspond to experimental data and the lines are fits. Vertical bars represent the Bragg reflections for the space group R3̄c. The difference pattern between the observed data and fits is shown at the bottom. The inset shows a zoom in the region between 2θ: 34–64°.
PMC9078421
c7ra13244a-f1.jpg
0.457102
5d6425f9ddc642c2aded19ac6c8b4446
(a) Crystal structure of La0.65Ce0.05Sr0.3MnO3 (b) 3D view showing MnO6 octahedron (c) SEM images of (x = 0 and x = 0.15) samples.
PMC9078421
c7ra13244a-f2.jpg
0.472865
c850d0a351704f59b6c32976a430ca33
Raman spectrum of La0.65Ce0.05Sr0.3Mn1−xCuxO3 (x = 0, 0.05, 0.10 and 0.15).
PMC9078421
c7ra13244a-f3.jpg
0.404806
015062f0308d4b01b5458738c38b49d1
Temperature dependence of the magnetization for La0.65Ce0.05Sr0.3Mn1−xCuxO3 (x = 0.10 and x = 0.15) measured in field cooling (FC) mode at an applied magnetic field of μ0H = 500 Oe. Inset (a) the temperature derivative dM/dT. (b) Temperature dependence of the inverse of magnetic susceptibility 1 = χ. The red line presents the linear fit at high temperature.
PMC9078421
c7ra13244a-f4.jpg
0.409873
db9bafc7ecd24ffdaeb3969a034006ad
Isothermal magnetization versus magnetic field around TC of La0.65Ce0.05Sr0.3Mn1−xCuxO3, (a) for x = 0 and (b) for x = 0.15.
PMC9078421
c7ra13244a-f5.jpg
0.469078
b141fac3c86b47c48cb8a5de3eef116d
Arrott plot of μ0H/M vs. M2 at different temperatures for La0.65Ce0.05Sr0.3Mn1−xCuxO3, (a) x = 0 and (b) x = 0.15.
PMC9078421
c7ra13244a-f6.jpg
0.475911
e82f19e2f95a4b799e2ad77fe226b4cb
The temperature dependence of the magnetic entropy change (ΔSM) under different applied magnetic fields (a) for x = 0, (b) for x = 0.05, (c) for x = 0.10 and (d) for x = 0.15.
PMC9078421
c7ra13244a-f7.jpg
0.470437
4e6d66b446f44910a0be8aeae2477515
Normalized ΔSMversus rescaled temperature θ for La0.65Ce0.05Sr0.3Mn1−xCuxO3, the solid line is the average curve.
PMC9078421
c7ra13244a-f8.jpg
0.396621
550080a883e948588c6d81f445fb9c96
Typical TEM images of PAH-MTX nanoassemblies: (A1) sample a1, (A2) sample a2, (A3) sample a3, (B1) sample b1, (B2) sample b2, (B3) sample b3, (C1) sample c1, (C2) sample c2, (C3) sample c3, (D1) sample d1, (D2) sample d2, (D3) sample d3.
PMC9078489
c7ra12862b-f1.jpg
0.415394
bbb3e87917544128980ead34f4a8ee95
Typical TEM images of PAH–MTX nanoassemblies: (E1) sample e1, (E2) sample e2, (E3) sample e3, (E4) sample e4, (F0) sample c1, (F1) sample f1, (F2) sample f2, (F3) sample f3.
PMC9078489
c7ra12862b-f2.jpg
0.42795
f018dc091ea64938ba2c871a553d101e
(A) FTIR spectra of free MTX, PAH, samples a2, b1, c2 and d1. (B) UV-vis absorption spectra of free MTX, PAH, samples a2, b1, c2 and d1. (C) XRD patterns of free MTX, PAH, samples a2, b1, c2 and d1.
PMC9078489
c7ra12862b-f3.jpg
0.445253
165ebd728829448b8a74251366f38ac6
(A) Release profiles of free MTX, samples a2, b1, c2 and d1. (B–E) Plots of different kinetic models for the release of MTX from the PAH–MTX nanoassemblies.
PMC9078489
c7ra12862b-f4.jpg
0.40863
f3498f7ad396475fb703ad61ed7f4845
(A) Comparison of cell viabilities for free MTX, samples a2, b1, c2 and d1 at various concentrations after 24 h of incubation; (B) comparison of cell viabilities for free MTX, sample a2, b1, c2 and d1 after different incubation time at the concentration of 100 μg mL−1. All cell viability data were obtained from three separated experiments and error bars represent standard error (n = 3).
PMC9078489
c7ra12862b-f5.jpg
0.505199
5b5ce355d0c3460392473193e84c4e5e
Morphology changes of A549 cells treated with various samples at the concentration of 100 μg mL−1: (A) DMEM for 24 h; (B) free MTX for 24 h; (C) sample a2 for 24 h; (D) sample a2 for 48 h; (E) sample a2 for 72 h.
PMC9078489
c7ra12862b-f6.jpg
0.527485
435d41a3b2194355878cf9c6957096bd
XRD patterns for the rock and soil samples ((A) <150 μm, (B) <2 μm). XRD revealed quartz (d = 4.25 and 3.34), feldspar (d = 3.76, 3.22, 2.98, and 2.90), and kaolinite (d = 7.15 and 3.55) as the possible main mineral constituents, while illite (d = 9.94 and 4.25) and montmorillonite (d = 15.29) were revealed as the possible minor mineral constituents in the rock and soil samples. Letter designations: Q, quartz; F, K-feldspar; K, kaolinite; I, illite; M, montmorillonite.
PMC9079951
c8ra01268g-f1.jpg
0.441272
4e731c48c37243bba939c0d3ab67590a
Influence of rock-weathering bacteria (L26, M23, and H41) on the releases of Fe, Si, and Al and on the pH of the medium added in tuff during 15 days of incubation. Error bars are ±standard error (n = 3).
PMC9079951
c8ra01268g-f2.jpg
0.466661
40f7d62929054e3b8056ea6361f96cc2
Influence of salt, temperature, and pH on the growth of the mineral-weathering bacteria isolated from the less and more altered rock samples and the soil samples. Error bars are ±standard error (n = 3). Bars indicated by the same letter are not significantly different (P > 0.05) according to Tukey’s test.
PMC9079951
c8ra01268g-f3.jpg
0.453233
4dbe93c318824e939799a57a576e73b5
Proportional abundance of the high, moderate, and low effective element (Si, Al, and Fe) solubilizers from the less (LR) and more (MR) altered rock and soil samples (SS), respectively.
PMC9079951
c8ra01268g-f4.jpg
0.428115
bade1bccaff34835b50c702d1b92b81a
(a) Spectroelectrochemistry of TTF1 in CH2Cl2 (c = 5 × 10−4 mol L−1); (b) UV-Vis absorption spectra and (c) ESR spectra of TTF1 and TTF2 upon adding 3 equivalents of I2 in CH2Cl2 (c = 1 × 10−5 mol L−1).
PMC9080441
c8ra02956c-f1.jpg
0.479121
c6f0cbb8354d47c0abcff86ebe87b611
Crystal structure of complex (TTF1)·(I3)·(I2): (a) top view of molecule TTF1 with the central CC bond length shown in unit of Å; (b) TTF1 dimer with atomic short contacts shown in dashed lines (green for S⋯S and grey for C⋯S); (c) anion sheets composed of (I3)− and I2 with the I–I bond length and I⋯I contacts (purple dashed lines) shown; (d) packing structure viewed along the longitudinal axis of TTF1 dimer with the I⋯I contacts shown in grey dashed lines.
PMC9080441
c8ra02956c-f2.jpg
0.465517
ddcb5a51b8964f87a4a59992b20474b2
Crystal structures of complex (TTF2)·(I5)·(I2): (a) top view of molecule TTF2 with the central CC bond length shown in unit of Å; (b) the (I3)− anion chain with the I–I bond lengths and I⋯I contacts (purple dashed lines) shown; (c) packing structure projected along the longitudinal axis of the TTF moiety.
PMC9080441
c8ra02956c-f3.jpg
0.467943
e145fd6269b24037bc56db0dfb609bc6
(a) ESR spectra for the crystalline complexes of (TTF1)·(I3)·(I2) and (TTF2)·(I5)·(I2); UV-Vis absorption spectra of (TTF1)·(I3)·(I2) and (TTF2)·(I5)·(I2) in the (b) solid state, and (c) CH2Cl2 solution (c = 10−5 mol L−1) after standing under inert atmosphere for 30 min and/or 24 h.
PMC9080441
c8ra02956c-f4.jpg