Journal articles on the topic 'Concentration cell'

To see the other types of publications on this topic, follow the link: Concentration cell.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Concentration cell.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Kusuma, Riska Anggri, Linda Suyati, and Wasino Hadi Rahmanto. "Effect of Lactose Concentration as Lactobacillus bulgaricus Substrate on Potential Cells Produced in Microbial Fuel Cell Systems." Jurnal Kimia Sains dan Aplikasi 21, no. 3 (July 31, 2018): 144–48. http://dx.doi.org/10.14710/jksa.21.3.144-148.

Full text
Abstract:
The effect of laxose concentration as Lactobacillus bulgaricus bacterial substrate on the cell potential produced in Microbial Fuel Cell System has been done. This study aims to determine the effect of lactose concentration as bacterial substrate, to generate electricity, maximum electric potential and determine the potential value of standard lactose (E ° Lactose.) Based on Nernst equation. The MFC system of two compartments and bridges of salt as a linkage is used in this study. Anode contains lactose with variation of concentration 3 - 7% and bacteria. The cathode contains a 1M KMO4. The electrodes used are graphite. MFC operational time is 14 days. The results showed that the lactose concentration had an effect on the cell potential produced in the MFC system. Maximum cell potential yielded at 4% lactose concentration, that is 710 mV then based on Nerst equation theory obtained E ° Lactose value in MFC system of + 0,236 V.
APA, Harvard, Vancouver, ISO, and other styles
2

Kisitu, Jaffar. "Chemical concentrations in cell culture compartments (C5) – concentration definitions." ALTEX 36, no. 1 (2019): 154–60. http://dx.doi.org/10.14573/altex.1901031.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Z, Ding. "Concentration Polarization of Ox-LDL and Its Effect on Cell Proliferation and Apoptosis in Human Endothelial Cells." Journal of Cardiology and Cardiovascular Medicine 1, no. 1 (2016): 011–18. http://dx.doi.org/10.29328/journal.jccm.1001003.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Kamath, Meghana, Isaac Houston, Alexander Janovski, Xiang Zhu, Sivakumar Gowrisankar, Anil Jegga, and Rodney DeKoter. "Myeloid Gene Activation and T Cell/Natural Killer Cell Gene Repression in Cells Expressing Two Distinct PU.1 Concentrations." Blood 110, no. 11 (November 16, 2007): 1242. http://dx.doi.org/10.1182/blood.v110.11.1242.1242.

Full text
Abstract:
Abstract The Ets transcription factor PU.1 (encoded by the gene Sfpi1) functions in a concentration-dependent manner as a hematopoietic cell fate determinant. PU.1 levels are uniform in early hematopoiesis, increase during myeloid differentiation, and decrease after erythrocyte and T cell/natural killer cell commitment. It is unknown how downstream target genes respond to changes in PU.1 concentration. To address this, we generated mice with two distinct hypomorphic alleles of Sfpi1 and analyzed interleukin-3 dependent cell lines from fetal liver cells homozygous for either allele. PU.1 was produced in these cells at ∼20% (Sfpi1BN/BN) or ∼2% (Sfpi1Blac/Blac) of wild type. These cells fail to terminally differentiate as a consequence of low PU.1 expression and can be maintained as cell lines. To determine what groups of genes are expressed in response to two distinct PU.1 concentrations, we performed whole-genome microarray analysis and compared gene expression in Sfpi1BN/BN and Sfpi1Blac/Blac cell lines to Sfpi1−/− cell lines. Groups of downstream target genes were activated or repressed in four modes in response to the two discrete concentrations of PU.1: at higher but not lower PU.1 concentration, at lower but not higher PU.1 concentration, at both lower and higher concentration, and in a gradient fashion. We decided to focus on genes regulated in a gradient manner, because dose-dependency suggests that these may be direct targets of PU.1. Genes activated in a gradient manner were mostly myeloid-specific and enriched for target genes of PU.1. Genes repressed in a gradient manner included erythroid-specific genes and, unexpectedly, T cell and natural killer cell-specific genes. T cell genes were also repressed by PU.1 in cultured progenitor-B cells. With this unique allelic system, we can study three discrete concentrations of PU.1 at 20%, 2%, and 0% to examine concentration-dependent effects of PU.1 on target genes and lineage decisions. Overall, our results suggest that PU.1 functions in a concentration-dependent manner to promote myeloid differentiation and repress T cell or natural killer cell development.
APA, Harvard, Vancouver, ISO, and other styles
5

Kočí, Vladimír, Darek Dragoun, and Jaromír Lukavský. "Determination of algal cell culture (Desmodesmus subspicatus) concentration using a microplate reader." Algological Studies/Archiv für Hydrobiologie, Supplement Volumes 122 (December 1, 2006): 123–35. http://dx.doi.org/10.1127/1864-1318/2006/0122-0123.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Rowley, SD, WI Bensinger, TA Gooley, and CD Buckner. "Effect of cell concentration on bone marrow and peripheral blood stem cell cryopreservation." Blood 83, no. 9 (May 1, 1994): 2731–36. http://dx.doi.org/10.1182/blood.v83.9.2731.2731.

Full text
Abstract:
Abstract The effects of cell concentration during cryopreservation on bone marrow (BM) or peripheral blood (PB)-derived hematopoietic progenitor cells have not been described. The much greater numbers of cells harvested for autologous PB stem cell (PBSC) transplantation requires that the cells be frozen at higher cell concentrations, or in much greater volumes, compared with BM. We cryopreserved 108 PBSC collections from 30 patients at an average (+/- SD) cell concentration of 3.7 +/- 1.9 x 10(8) nucleated cells per mL in 127 +/- 45 mL. The proportion of mononuclear cells was 52.9% +/- 27.2%. The products also contained 2.9 +/- 2.1 x 10(9) platelets/mL and an average red cell proportion of 12.9% +/- 7.2%. The nucleated cell recovery after thawing was 75.4% +/- 13.0%. The nucleated cell concentration during freezing was not predictive for the postthaw recoveries of nucleated cells (P = .38), granulocyte-macrophage colony-forming unit (P = .06) or CD34+ cells (P = .54), or for the viability of mononuclear cells (P = .81). The platelet and red cell concentrations similarly were not predictive for these endpoints. Samples (3 BM, 7 PBSC) from 10 patients were simultaneously cryopreserved at two-fold, and from 5 additional patients (PBSC) at 6- to 24-fold differing cell concentrations. A lower recovery of erythroid burst forming unit was found for samples frozen at higher cell concentrations (P = .04), but no significant differences were found in the other endpoints listed above. The average cell concentration during freezing for each patient's PBSC collections (n = 34 patients) did not predict time to achieve a PB count of > 500 granulocytes/microL (P = .51) or platelet transfusion independence (P = .39). Patients achieved these endpoints of engraftment at medians of 12 and 13 days, respectively. The infusion of these products was generally well tolerated. Similarly, the cell concentration at which BM cells were frozen did not predict for the duration of granulocyte (P = .63) or platelet (P = .36) aplasias for 54 patients undergoing autologous BM transplantation. These data suggest that PBSC or BM cells collected for transplantation may be cryopreserved at very high cell concentrations without loss of engraftment potential or undue infusion-related toxicity.
APA, Harvard, Vancouver, ISO, and other styles
7

Rowley, SD, WI Bensinger, TA Gooley, and CD Buckner. "Effect of cell concentration on bone marrow and peripheral blood stem cell cryopreservation." Blood 83, no. 9 (May 1, 1994): 2731–36. http://dx.doi.org/10.1182/blood.v83.9.2731.bloodjournal8392731.

Full text
Abstract:
The effects of cell concentration during cryopreservation on bone marrow (BM) or peripheral blood (PB)-derived hematopoietic progenitor cells have not been described. The much greater numbers of cells harvested for autologous PB stem cell (PBSC) transplantation requires that the cells be frozen at higher cell concentrations, or in much greater volumes, compared with BM. We cryopreserved 108 PBSC collections from 30 patients at an average (+/- SD) cell concentration of 3.7 +/- 1.9 x 10(8) nucleated cells per mL in 127 +/- 45 mL. The proportion of mononuclear cells was 52.9% +/- 27.2%. The products also contained 2.9 +/- 2.1 x 10(9) platelets/mL and an average red cell proportion of 12.9% +/- 7.2%. The nucleated cell recovery after thawing was 75.4% +/- 13.0%. The nucleated cell concentration during freezing was not predictive for the postthaw recoveries of nucleated cells (P = .38), granulocyte-macrophage colony-forming unit (P = .06) or CD34+ cells (P = .54), or for the viability of mononuclear cells (P = .81). The platelet and red cell concentrations similarly were not predictive for these endpoints. Samples (3 BM, 7 PBSC) from 10 patients were simultaneously cryopreserved at two-fold, and from 5 additional patients (PBSC) at 6- to 24-fold differing cell concentrations. A lower recovery of erythroid burst forming unit was found for samples frozen at higher cell concentrations (P = .04), but no significant differences were found in the other endpoints listed above. The average cell concentration during freezing for each patient's PBSC collections (n = 34 patients) did not predict time to achieve a PB count of > 500 granulocytes/microL (P = .51) or platelet transfusion independence (P = .39). Patients achieved these endpoints of engraftment at medians of 12 and 13 days, respectively. The infusion of these products was generally well tolerated. Similarly, the cell concentration at which BM cells were frozen did not predict for the duration of granulocyte (P = .63) or platelet (P = .36) aplasias for 54 patients undergoing autologous BM transplantation. These data suggest that PBSC or BM cells collected for transplantation may be cryopreserved at very high cell concentrations without loss of engraftment potential or undue infusion-related toxicity.
APA, Harvard, Vancouver, ISO, and other styles
8

Gülden, M., S. Mörchel, and H. Seibert. "Factors influencing nominal effective concentrations of chemical compounds in vitro: cell concentration." Toxicology in Vitro 15, no. 3 (June 2001): 233–43. http://dx.doi.org/10.1016/s0887-2333(01)00008-x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Trifilio, Steven M., Paul R. Yarnold, Marc H. Scheetz, Judy Pi, Gennethel Pennick, and Jayesh Mehta. "Serial Plasma Voriconazole Concentrations after Allogeneic Hematopoietic Stem Cell Transplantation." Antimicrobial Agents and Chemotherapy 53, no. 5 (February 17, 2009): 1793–96. http://dx.doi.org/10.1128/aac.01316-08.

Full text
Abstract:
ABSTRACT Plasma voriconazole concentrations vary considerably between patients receiving standard dosing, and trough voriconazole concentrations are known to affect efficacy and toxicity. Temporal variations in serial plasma voriconazole concentrations through the course of therapy in hematopoietic stem cell transplantation patients has not been carefully described. Paired voriconazole concentrations in 64 patients were studied to determine the predictability of the second concentration based on the first. The difference between the two values was ≤5% in six patients. In 25 patients, the second concentration was higher by a median of 40%. In 33 patients, the subsequent concentration was lower by a median of 59%. For patients with an initial concentration of <2 μg/ml, the correlation between the two values was poor (r = 0.24; P < 0.17). For those with an initial concentration of ≥2 μg/ml, the correlation was good (r = 0.72; P < 0.0001). There was no relationship between the magnitude of the change and the time elapsing between the two measurements. Among the 43 patients who had an initial concentration of ≥1 μg/ml, the two voriconazole measurements were strongly correlated (r = 0.66, P < 0.0001), but only 67% had a voriconazole serum concentration of ≥1 μg/ml on the second measurement. No studied variables were reliable predictors in identifying concentrations above or below 1 or 2 μg/ml. Our data suggest that variations in voriconazole concentrations are unpredictable despite standard dosing, and the acceptability of a concentration on one occasion cannot be extrapolated to future concentrations in the same patient. This suggests that ongoing therapeutic drug monitoring and dose adjustment may be beneficial in patients requiring prolonged voriconazole therapy.
APA, Harvard, Vancouver, ISO, and other styles
10

Jia, Chen, Abhyudai Singh, and Ramon Grima. "Concentration fluctuations in growing and dividing cells: Insights into the emergence of concentration homeostasis." PLOS Computational Biology 18, no. 10 (October 4, 2022): e1010574. http://dx.doi.org/10.1371/journal.pcbi.1010574.

Full text
Abstract:
Intracellular reaction rates depend on concentrations and hence their levels are often regulated. However classical models of stochastic gene expression lack a cell size description and cannot be used to predict noise in concentrations. Here, we construct a model of gene product dynamics that includes a description of cell growth, cell division, size-dependent gene expression, gene dosage compensation, and size control mechanisms that can vary with the cell cycle phase. We obtain expressions for the approximate distributions and power spectra of concentration fluctuations which lead to insight into the emergence of concentration homeostasis. We find that (i) the conditions necessary to suppress cell division-induced concentration oscillations are difficult to achieve; (ii) mRNA concentration and number distributions can have different number of modes; (iii) two-layer size control strategies such as sizer-timer or adder-timer are ideal because they maintain constant mean concentrations whilst minimising concentration noise; (iv) accurate concentration homeostasis requires a fine tuning of dosage compensation, replication timing, and size-dependent gene expression; (v) deviations from perfect concentration homeostasis show up as deviations of the concentration distribution from a gamma distribution. Some of these predictions are confirmed using data for E. coli, fission yeast, and budding yeast.
APA, Harvard, Vancouver, ISO, and other styles
11

Katsogiannou, EG, PD Katsoulos, C. Ziogas, MC Naskou, G. Christodoulopoulos, ZS Polizopoulou, A. Tzivara, and LV Athanasiou. "Blood cell count and morphology, and vitamin B12 concentration in pre- and post-weaned calves." Veterinární Medicína 66, No. 12 (October 26, 2021): 513–19. http://dx.doi.org/10.17221/12/2021-vetmed.

Full text
Abstract:
Haematological indicators may resent physiological variation by age. Vitamin B12 promotes haematopoiesis. The aims of this study were: 1) to compare the values of the haematological variables and the concentration of vitamin B12 in pre- or post-weaned veal calves and 2) to identify the possible association between the values of the haematological variables and the concentration of B12 in the blood of veal calves. Blood was collected on the same day from 31 pre-weaned and 31 weaned calves of the Limousine breed from the same farm. The complete blood count, including the blood cell morphology evaluation, was performed and the serum B12, total protein and albumin concentrations were determined. The serum concentration of vitamin B12, the haematocrit (HCT), the haemoglobin concentration (HGB), the platelet count and the lymphocyte count were significantly higher in the weaned calves. A very strong positive correlation was found between the concentration of the vitamin B12 and HCT and HGB before weaning, while these correlations were moderately positive following weaning and in the total population tested as well. The observed variation in the blood cell count and morphology, such as poikilocytosis and the presence of macrocytes and hypersegmented neutrophils, along with the age of the animal seem to be related to the vitamin B12 concentration.
APA, Harvard, Vancouver, ISO, and other styles
12

Dinh Lam, Nguyen, Le Thuy Trang, Nguyen Thi Mui, Pham Van Vinh, Vuong Van Cuong, and Nguyen Van Hung. "INFLUENCES OF Sn DOPING CONCENTRATION ON CHARACTERISTICS OF ZnO FILMS FOR SOLAR CELL APPLICATIONS." Journal of Science, Natural Science 60, no. 7 (2015): 41–46. http://dx.doi.org/10.18173/2354-1059.2015-0030.

Full text
APA, Harvard, Vancouver, ISO, and other styles
13

Kiuchi, Tai, Tomoaki Nagai, Kazumasa Ohashi, and Kensaku Mizuno. "Measurements of spatiotemporal changes in G-actin concentration reveal its effect on stimulus-induced actin assembly and lamellipodium extension." Journal of Cell Biology 193, no. 2 (April 18, 2011): 365–80. http://dx.doi.org/10.1083/jcb.201101035.

Full text
Abstract:
To understand the intracellular role of G-actin concentration in stimulus-induced actin assembly and lamellipodium extension during cell migration, we developed a novel technique for quantifying spatiotemporal changes in G-actin concentration in live cells, consisting of sequential measurements of fluorescent decay after photoactivation (FDAP) of Dronpa-labeled actin. Cytoplasmic G-actin concentrations decreased by ∼40% immediately after cell stimulation and thereafter the cell area extended. The extent of stimulus-induced G-actin loss and cell extension correlated linearly with G-actin concentration in unstimulated cells, even at concentrations much higher than the critical concentration of actin filaments, indicating that cytoplasmic G-actin concentration is a critical parameter for determining the extent of stimulus-induced G-actin assembly and cell extension. Multipoint FDAP analysis revealed that G-actin concentration in lamellipodia was comparable to that in the cell body. We also assessed the cellular concentrations of free G-actin, profilin- and thymosin-β4–bound G-actin, and free barbed and pointed ends of actin filaments by model fitting of jasplakinolide-induced temporal changes in G-actin concentration.
APA, Harvard, Vancouver, ISO, and other styles
14

Han, Song-I., Hyun Soo Kim, and Arum Han. "In-droplet cell concentration using dielectrophoresis." Biosensors and Bioelectronics 97 (November 2017): 41–45. http://dx.doi.org/10.1016/j.bios.2017.05.036.

Full text
APA, Harvard, Vancouver, ISO, and other styles
15

Kilbride, Peter, Julie Meneghel, and John Morris. "Cell Concentration And Sedimentation In Cryopreservation." Cryobiology 91 (December 2019): 193. http://dx.doi.org/10.1016/j.cryobiol.2019.10.180.

Full text
APA, Harvard, Vancouver, ISO, and other styles
16

Hassan, Adnan F., Ibtihaj R. Alshirmani, and Ali H. khwayyir. "Pyranine Dye as Solar Cell Concentration." International Journal of Applied Physics 6, no. 2 (May 25, 2019): 73–78. http://dx.doi.org/10.14445/23500301/ijap-v6i2p111.

Full text
APA, Harvard, Vancouver, ISO, and other styles
17

Loges, K. M., P. Wiedemann, B. Hitzmann, and K. Preuß. "Automated cell concentration control in bioreactors." Chemie Ingenieur Technik 92, no. 9 (August 28, 2020): 1337–38. http://dx.doi.org/10.1002/cite.202055091.

Full text
APA, Harvard, Vancouver, ISO, and other styles
18

Shah, Syed Mohmad, Neha Saini, Syma Ashraf, Manoj Kumar Singh, Radhey Sham Manik, Suresh Kumar Singla, Prabhat Palta, and Manmohan Singh Chauhan. "Cumulus cell-conditioned medium supports embryonic stem cell differentiation to germ cell-like cells." Reproduction, Fertility and Development 29, no. 4 (2017): 679. http://dx.doi.org/10.1071/rd15159.

Full text
Abstract:
Cumulus cells provide cellular interactions and growth factors required for oogenesis. In vitro studies of oogenesis are limited primarily because of the paucity of their source, first trimester fetal gonads, and the small number of germ lineage precursor cells present within these tissues. In order to understand this obscure but vitally important process, the present study was designed to direct differentiation of embryonic stem (ES) cells into germ lineage cells. For this purpose, buffalo ES cells were differentiated, as embryoid bodies (EBs) and monolayer adherent cultures, in the presence of different concentrations of cumulus-conditioned medium (CCM; 10%, 20% and 40%) for different periods of culture (4, 8 and 14 days) to identify the optimum differentiation-inducing concentration and time. Quantitative polymerase chain reaction analysis revealed that 20%–40% CCM induced the highest expression of primordial germ cell-specific (deleted in Azoospermia- like (Dazl), dead (Asp-Glu-Ala-Asp) box polypeptide 4 (Vasa also known as DDX4) and promyelocytic leukemia zinc finger protein (Plzf)); meiotic (synaptonemal complex protein 3 (Sycp3), mutl homolog I (Mlh1), transition protein 1/2 (Tnp1/2) and protamine 2 (Prm2); spermatocyte-specific boule-like RNA binding protein (Boule) and tektin 1 (Tekt1)) and oocyte-specific growth differentiation factor 9 (Gdf9) and zona pellucida 2 /3 (Zp2/3)) genes over 8–14 days in culture. Immunocytochemical analysis revealed expression of primordial germ cell (c-KIT, DAZL and VASA), meiotic (SYCP3, MLH1 and PROTAMINE 1), spermatocyte (ACROSIN and HAPRIN) and oocyte (GDF9 and ZP4) markers in both EBs and monolayer differentiation cultures. Western blotting revealed germ lineage-specific protein expression in Day 14 EBs. The significantly lower (P < 0.05) concentration of 5-methyl-2-deoxycytidine in differentiated EBs compared to undifferentiated EBs suggests that methylation erasure may have occurred. Oocyte-like structures obtained in monolayer differentiation stained positive for ZONA PELLUCIDA protein 4 and progressed through various embryo-like developmental stages in extended cultures.
APA, Harvard, Vancouver, ISO, and other styles
19

Feng, Zhenyu, Chao Chen, Zhenyu Ye, Jiaming Xie, Wei Chen, Wei Li, and Chungen Xing. "The Effect of Icariin on Human Hepatoellular Carcinomas Cell Apoptosis and Cell Cycle Arrest." Journal of Biomaterials and Tissue Engineering 11, no. 12 (December 1, 2021): 2389–94. http://dx.doi.org/10.1166/jbt.2021.2814.

Full text
Abstract:
Background: Natural products often have novel frameworks and unique mechanisms, It is an important way to develop new anti-tumor drugs.The paper explored the effect of Icariin on HepG2 cell apoptosis and cell cycle arrest. Material and methods: Human liver cancer HepG2 cells were studied. The biological activity of Icariin on HepG2 cells was comprehensively investigated. The mechanism was preliminarily explored from the aspects of proliferation, cell cycle, and apoptosis. Results: Icariin showed significant inhibitory effect on cell proliferation after administration, and the effect was time-and-concentration-dependent. Annexin V-PI detection showed that, after 48 hours of administration of Icariin at different concentrations, the apoptosis rate of HepG2 cells increased in a concentration-dependent manner. The results of Hoechst 33342 staining showed that, after 48 hours of intervention with Icariin at different concentrations, HepG2 cells appeared densely stained and granular fluorescence, characterizing apoptosis. In the JC-1 mitochondrial membrane potential experiment, Icariin was found to destroy cell mitochondrial membrane potential and induce HepG2 cell apoptosis. After 48 h administration of Icariin at different concentrations, Bcl-2 and survivin proteins were down-regulated, and Bax was up-regulated, both in a concentration-dependent manner. PI single staining combined with flow cytometry to detect cell cycle results showed that, Icariin can induce G2/M phase arrest of HepG2 cells, and is time-and-concentration-dependent. Western Blot detection revealed that Icariin can down-regulate the cycle-related proteins Cyclin B and CDK1 in a concentration-dependent manner, and also significantly down-regulate the expressions of p-AKT, AKT, p-ERK and ERK proteins. Conclusion: Icariin is a selectively potential active compound that can treat liver cancer. The paper provided theoretical basis and experimental support for the clinical application of Icariin in the drug treatment of liver cancer. It is necessary to further study the antitumor effect of Icariin, explore and find the target, and provide higher selectivity for the treatment of liver cancer by Icariin.
APA, Harvard, Vancouver, ISO, and other styles
20

Johansen, P. W., O. P. F. Clausen, E. Haug, S. Fossum, and K. M. Gautvik. "Effects of bromocriptine on cell cycle distribution and cell morphology in cultured rat pituitary adenoma cells." Acta Endocrinologica 110, no. 3 (November 1985): 319–28. http://dx.doi.org/10.1530/acta.0.1100319.

Full text
Abstract:
Abstract. The effects of bromocriptine, a dopamine (DA) agonist, on cell cycle distribution and cell morphology have been studied in a clonal strain of rat pituitary adenoma cells (GH3) which produce and secrete spontaneously both prolactin (Prl) and growth hormone (GH). DNA flow cytometry showed that bromocriptine caused a dose-dependent delay in cell cycle traverse concomitantly with a reduction in cellular growth rate. The lowest concentration of bromocriptine (5 × 10−6 mol/l) significantly (P < 0.05) increased the relative number of cells in the S phase and reduced the proportion of cells in the G1 phase. At higher concentrations (1 × 10−5–5 × 10−5 mol/l) bromocriptine delayed cell cycle traverse through effects on cells in the S, G1 and G2 phases. These effects occurred already after 24 h of treatment. These results were supported by autoradiography of nuclear uptake of [3H]thymidine and by measurements of the number of cells arrested in metaphase after colcemide treatment (mitotic rate). Bromocriptine at 5 × 10−5 mol/l altered profoundly GH3 cell structure inducing cell clustering and typical changes in mitochondrial and nuclear ultrastructures. Since Prl and GH production is a characteristic of cells in G1 phase, the inhibitory effect of the lowest antiproliferative concentration of bromocriptine (5 × 10−6 mol/l) can only partly be explained by alterations in phase distribution. At the highest concentration of bromocriptine (5 × 10−5 mol/l) hormone production and cell division are also inhibited due to general toxic effects as reflected by the ultrastructural changes.
APA, Harvard, Vancouver, ISO, and other styles
21

Wang, Zilong, Hua Zhang, Wei Zhao, Zhigang Zhou, and Mengxun Chen. "The Effect of Concentrated Light Intensity on Temperature Coefficient of the InGaP/InGaAs/Ge Triple-Junction Solar Cell." Open Fuels & Energy Science Journal 8, no. 1 (May 29, 2015): 106–11. http://dx.doi.org/10.2174/1876973x01508010106.

Full text
Abstract:
Research on automatic tracking solar concentrator photovoltaic systems has gained increasing attention in developing the solar PV technology. A paraboloidal concentrator with secondary optic is developed for a three-junction GaInP/GalnAs/Ge solar cell. The concentration ratio of this system is 200 and the photovoltaic cell is cooled by the heat pipe. A detailed analysis on the temperature coefficient influence factors of triple-junction solar cell under different high concentrations (75X, 100X, 125X, 150X, 175X and 200X) has been conducted based on the dish-style concentration photovoltaic system. The results show that under high concentrated light intensity, the temperature coefficient of Voc of triple-junction solar cell is increasing as the concentration ratio increases, from -10.84 mV/°C @ 75X growth to -4.73mV/°C @ 200X. At low concentration, the temperature coefficient of Voc increases rapidly, and then increases slowly as the concentration ratio increases. The temperature dependence of η increased from -0.346%/°C @ 75X growth to - 0.103%/°C @ 200X and the temperature dependence of Pmm and FF increased from -0.125 W/°C, -0.35%/°C @ 75X growth to -0.048W/°C, -0.076%/°C @ 200X respectively. It indicated that the temperature coefficient of three-junction GaInP/GalnAs/Ge solar cell is better than that of crystalline silicon cell array under concentrating light intensity.
APA, Harvard, Vancouver, ISO, and other styles
22

Chow, S. Y., Y. C. Yen-Chow, and D. M. Woodbury. "Water and electrolyte contents, cell pH and membrane potentials of cultured turtle thyroid cells." Journal of Endocrinology 104, no. 1 (January 1985): 45–52. http://dx.doi.org/10.1677/joe.0.1040045.

Full text
Abstract:
ABSTRACT Water and electrolyte contents, cell pH, membrane potential and 125I− uptake were determined in cultured follicular cells of turtle thyroid. The Na+, K+ and Cl− concentrations in the cultured thyroid cells were 59·2, 119·0 and 50·9 mmol/l cell water respectively. Treatment with TSH (10 mu./ml for 24 h) increased the K+ and Cl− and decreased the Na+ concentrations in cells. The water and protein contents of these cells were 81·6 and 8·7 g/100 g cells respectively. The cell pH was 6·91. With glass microelectrodes, the resting membrane potential of thyroid cells cultured in Medium 199 averaged 33·9 ± 0·63 mV which is slightly higher than 29·8 ± 1·6 mV as calculated from the data on the uptakes of [14C]methyltriphenylphosphonium and 3H2O by the cells. The potential varied linearly with the log of external K + concentration (between 15 and 120 mmol/l) with a slope of about 24 mV per tenfold change in K+ concentration. Both TSH and cyclic AMP depolarized the cell membrane. Calculations based on the values for the electrolyte concentrations in cells and in culture medium indicated that Na+, K+ and Cl− were not distributed according to their electrochemical gradients across the cell membrane. Na+ was actively transported out of the cells and K+ and Cl− into the cells. Follicular cells of turtle thyroid cultured in the medium without addition of TSH formed a monolayer. Their iodide-concentrating ability was low and they did not respond to TSH with an increase in iodide uptake. In contrast, cells cultured in medium containing TSH tended to aggregate and organize to form follicles. They had higher ability to concentrate iodide and respond to TSH. J. Endocr. (1985) 104, 45–52
APA, Harvard, Vancouver, ISO, and other styles
23

Jiang, Guojun, Jiahua Hu, Junhe Huang, Yan Li, Qing Deng, Wenyan Jiang, Guihong Huang, and Qingqing Wang. "The Effect of Isoquercitrin on Cell Apoptosis and Cycle for HepG2 Cells." Scholars Academic Journal of Pharmacy 11, no. 10 (November 29, 2022): 182–86. http://dx.doi.org/10.36347/sajp.2022.v11i10.002.

Full text
Abstract:
Purpose: To study the effect of isoquercitrin on cell apoptosis and cycle of HepG2 cells. Materials and Method: Different concentrations of isoquercitrin were used to act on HepG2 cells, and the effect of isoquercitrin on the proliferation of HepG2 cells was detected by CCK8. Cell morphology and growth were observed under inverted microscope. Flow cytometry was used to detect the apoptosis and cycle changes of HepG2 cells. Results: CCK8 found that isoquercitrin inhibited the growth of HepG2 cells, and it was correlated with the concentration and action time of isoquercitrin. Under the inverted microscope, it was observed that the number of cell survival gradually decreased with the increase of concentration or time of isoquercitrin acted on HepG2 cells. Flow cytometry showed that with the increase of isoquercitrin concentration, the number of cells blocked in S phase gradually decreased and the number of cells blocked in G2/M phase gradually increased. Conclusion: Isoquercitrin can induce apoptosis of HepG2 cells and interfere with S phase and G2/M phase in cell cycle.
APA, Harvard, Vancouver, ISO, and other styles
24

Liu, Sheng-Long, Lu Yang, Cheng-Jun Zhu, Kai Liu, Wei Han, and Jia-Feng Yao. "A method of identifying cell suspension concentration based on bioimpedance spectroscopy." Acta Physica Sinica 71, no. 7 (2022): 078701. http://dx.doi.org/10.7498/aps.71.20211837.

Full text
Abstract:
Based on bioimpedance spectroscopy technology, a method of automatically identifying the cell suspension concentration is proposed. This method combines multiple linear regression algorithm and bioimpedance spectroscopy technology, which can identify the concentration of cell suspension quickly and accurately. Firstly, a strategy of random distribution of cell locations is proposed to simulate the true existence of cells. Secondly, 2400 groups of normal, cancerous and mixed cell models with different concentrations are generated by numerical simulation and their bioimpedance spectroscopy data are calculated.Thirdly, the multiple linear regression algorithm (MLR), support vector machine (SVM), and gradient boosting regression algorithm (GBR) are used to identify the concentration of cancerous cells. The simulation results show that the MLR is the best regression model for cell suspension concentration identification and its average goodness of fit and mean square error are 0.9997 and 0.0008respectively. Finally, the MLR is applied to the identification of red blood cell suspensions with different concentrations, the experimental results show that the average goodness of fit and mean square error are 0.9998 and 0.0079, respectively, indicating that this method has a greater ability to identify cell suspension concentrations.
APA, Harvard, Vancouver, ISO, and other styles
25

Salmman, Israa Sekar. "Effect crude aquatic extract of leaves Ocimum basilicum on caner and normal cell lines in vitro." Journal of Biotechnology Research Center 7, no. 1 (January 1, 2013): 5–13. http://dx.doi.org/10.24126/jobrc.2013.7.1.233.

Full text
Abstract:
Leaves of Ocimum basilicum were extracted with distilled water to prepare aquatic crude extract, five concentration from this extract (1000, 500, 250, 125, 62.5)µg/ml were used to study the effect of extract on cancer cell lines(Larynex carcinoma hep-2, cervix carcinoma Hela and mammary gland adenocarcinoma AMN-3) and the time exposure 24 and 48 hours as well as the effect of aquous extract on normal cell line for embryonic mice fibroblast(MEF) was studied in vitro at 48hrs. exposure. The result showed that aqueous crude extract of Ocimum basilicum leaves has different effects on cancer cell lines with significante p<0.05 the high concentration 1000 µg/ml has more inhibitory effect on cancer cell line Hep-2 compared with low concentration at 24and 48 hrs.while the Hela cancer cell line has hormetic effect which recognized by contrast in low concentration inhibitory effect as compared with high concentration at 24 and 48 hr.that low ones inhibit cell proliferation while in high ones cell proliferation continue, but AMN-3 cancer cell line more affected by low concentrations from high concentrations. Normal cell line show no significant effect for all concentrations used of aquatic crude extract of Ocimum basilicum leaf except 62.5 µg/ml with high cell inhibition 16%.
APA, Harvard, Vancouver, ISO, and other styles
26

Wagner, Johanna E., Janice L. Huff, William L. Rust, Karl Kingsley, and George E. Plopper. "Perillyl Alcohol Inhibits Breast Cell Migration without Affecting Cell Adhesion." Journal of Biomedicine and Biotechnology 2, no. 3 (2002): 136–40. http://dx.doi.org/10.1155/s1110724302207020.

Full text
Abstract:
The monoterpene d-limonene exhibits chemotherapeutic and chemopreventive potential in breast cancer patients. D-limonene and its related compounds, perillyl alcohol and perillyl aldehyde, were chosen as candidate drugs for application in a screen for nontoxic inhibitors of cell migration. Using the nontumorigenic human breast cell line MCF-10A, we delineated the toxicity as greatest for the perillyl aldehyde, intermediate for perillyl alcohol, and least for limonene. A noncytotoxic concentration of 0.5 mmol/L perillyl alcohol inhibited the migration, while the same concentration of limonene failed to do so. Adhesion of the MCF-10A cell line and the human breast cancer cell line MDA-MB 435 to fibronectin was unaffected by 1.5 mmol/L perillyl alcohol. 0.4 mmol/L perillyl alcohol inhibited the growth of MDA-MB 435 cells. All migration-inhibiting concentrations of perillyl alcohol for MDA-MB 435 cells proved to be toxic. These results suggest that subtoxic doses of perillyl alcohol may have prophylactic potential in the treatment of breast cancer.
APA, Harvard, Vancouver, ISO, and other styles
27

Brignardello, E., M. Gallo, M. Aragno, R. Manti, E. Tamagno, O. Danni, and G. Boccuzzi. "Dehydroepiandrosterone prevents lipid peroxidation and cell growth inhibition induced by high glucose concentration in cultured rat mesangial cells." Journal of Endocrinology 166, no. 2 (August 1, 2000): 401–6. http://dx.doi.org/10.1677/joe.0.1660401.

Full text
Abstract:
The oxidative stress induced by high glucose concentration contributes to tissue damage associated with diabetes, including renal injury. Dehydroepiandrosterone (DHEA), the major secretory product of the human adrenal gland, has been shown to possess a multi-targeted antioxidant activity which is also effective against lipid peroxidation induced by high glucose. In this study we evaluated the effect of DHEA on the growth impairment which high glucose concentration induces in cultured rat mesangial cells. Primary cultures of rat mesangial cells were grown for 10 days in media containing either normal (i.e. 5.6 mmol/l) or high (i.e. 30 mmol/l) concentrations of glucose, without or with DHEA at different concentrations. The impairment of cell growth induced by high glucose was reversed by 100 nmol/l and 500 nmol/l DHEA, which had no effect on mesangial cells cultured in media containing glucose at the normal physiological concentration (5.6 mmol/l). In high-glucose cultured mesangial cells, DHEA also attenuated the lipid peroxidation, as measured by thiobarbituric acid reactive substances (TBARS) generation and 4-hydroxynonenal (HNE) concentration, and preserved the cellular content of reduced glutathione as well as the membrane Na+/K+ ATPase activity. The data further support the protective effect of DHEA against oxidative damage induced by high glucose concentrations, and bring into focus its possible effectiveness in preventing chronic complications of diabetes.
APA, Harvard, Vancouver, ISO, and other styles
28

Liu, Tianbin, Jie Li, Hongwei Dou, Xi Xiang, Wenbin Chen, Tingting Zhang, Lin Li, et al. "Low-Concentration Essential Amino Acids in PZM-3 Improve the Developmental Competence of Porcine Embryos Produced by Handmade Cloning." Cellular Reprogramming 22, no. 6 (December 1, 2020): 282–90. http://dx.doi.org/10.1089/cell.2020.0036.

Full text
APA, Harvard, Vancouver, ISO, and other styles
29

NISHIBAYASHI, Kento, Daisuke KAWASHIMA, Liu XIAYI, Hiromichi OBARA, and Masahiro TAKEI. "Cell concentration and dead cell rate sensing of yeast cell by impedance measurement." Proceedings of the Bioengineering Conference Annual Meeting of BED/JSME 2019.32 (2019): 1C33. http://dx.doi.org/10.1299/jsmebio.2019.32.1c33.

Full text
APA, Harvard, Vancouver, ISO, and other styles
30

He, Lihong, Xiaorui Wang, Qing Ma, Weipeng Zhao, Yongsheng Jia, Guolei Dong, Yuehong Zhu, Xiaochen Jia, and Zhongsheng Tong. "Ginsenoside induces cell death in breast cancer cells via ROS/PI3K/Akt signaling pathway." Tropical Journal of Pharmaceutical Research 19, no. 8 (November 18, 2020): 1631–36. http://dx.doi.org/10.4314/tjpr.v19i8.10.

Full text
Abstract:
Purpose: To study the influence of ginsenoside on breast carcinoma, and the mechanism of action involved.Methods: Different concentrations of ginsenoside were used to treat MCF-7 breast cancer cell line. Cell viability was measured by MTT assay, while protein expressions of p-Akt and p-PI3K were determined using Western blotting. The concentrations of reactive oxidative reactants and reactive oxygen species (ROS) were assessed using fluorescence immunoassay and immunofluorescence assay. The mechanism of action involved in ginsenoside-mediated apoptosis was determined based on ROS/PI3K/Akt signaling pathway.Results: There was no change in the inhibition of MCF-7 cell proliferation in control cells with time (p > 0.05). However, inhibition of MCF-7 cell proliferation in ginsenoside group was significantly higher than that in the control group (p < 0.05); furthermore, it increased with time and ginsenoside concentration. Apoptosis was markedly and concentration-dependently higher in ginsenoside-treated MCF-7 cells than in controls (p > 0.05). There were lower protein levels of p-PI3K and p-Akt in ginsenoside-exposed MCF-7 cells than in control group; the protein expressions decreased with increase in ginsenoside concentration (p < 0.05). The expressions of ROS in ginsenoside-treated MCF-7 cells declined, relative to the untreated group; in addition, the expressions decreased with increase in ginsenoside concentration (p < 0.05).Conclusion: Ginsenoside suppresses proliferation of MCF-7 cell line, and exerts apoptotic effect on the cells via inhibition of the ROS/PI3K/Akt signal pathway. This provides a new approach to treat breast cancer. Keywords: Breast cancer cells, Ginsenoside, Apoptosis, ROS/PI3K/Akt signaling pathway
APA, Harvard, Vancouver, ISO, and other styles
31

Li, Anqi, Gina D. Kusuma, Dawn Driscoll, Nathan Smith, Dominic M. Wall, Bruce L. Levine, David James, and Rebecca Lim. "Advances in automated cell washing and concentration." Cytotherapy 23, no. 9 (September 2021): 774–86. http://dx.doi.org/10.1016/j.jcyt.2021.04.003.

Full text
APA, Harvard, Vancouver, ISO, and other styles
32

Suehiro, J., R. Yatsunami, R. Hamada, and M. Hara. "Cell concentration determination using dielectrophoretic impedance measurement." Seibutsu Butsuri 39, supplement (1999): S205. http://dx.doi.org/10.2142/biophys.39.s205_3.

Full text
APA, Harvard, Vancouver, ISO, and other styles
33

Sumesh, C. G., A. V. Kumar, R. N. Nair, R. M. Tripathi, and V. D. Puranik. "Estimation of thoron concentration using scintillation cell." Radiation Protection Dosimetry 150, no. 4 (January 5, 2012): 536–40. http://dx.doi.org/10.1093/rpd/ncr438.

Full text
APA, Harvard, Vancouver, ISO, and other styles
34

CLARK, MARGARET R. "Mean Corpuscular Hemoglobin Concentration and Cell Deformability." Annals of the New York Academy of Sciences 565, no. 1 Sickle Cell D (July 1989): 284–94. http://dx.doi.org/10.1111/j.1749-6632.1989.tb24176.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
35

Cookson, Natalie A., Scott W. Cookson, Lev S. Tsimring, and Jeff Hasty. "Cell cycle-dependent variations in protein concentration." Nucleic Acids Research 38, no. 8 (December 17, 2009): 2676–81. http://dx.doi.org/10.1093/nar/gkp1069.

Full text
APA, Harvard, Vancouver, ISO, and other styles
36

Barnes, Robert A., Alan R. Bandy, and Arnold L. Torres. "Electrochemical concentration cell ozonesonde accuracy and precision." Journal of Geophysical Research: Atmospheres 90, no. D5 (August 20, 1985): 7881–87. http://dx.doi.org/10.1029/jd090id05p07881.

Full text
APA, Harvard, Vancouver, ISO, and other styles
37

Wigham, C. G., M. Williams, and S. A. Hodson. "Rabbit corneal endothelial cell intracellular sodium concentration." Experimental Eye Research 55 (September 1992): 104. http://dx.doi.org/10.1016/0014-4835(92)90566-b.

Full text
APA, Harvard, Vancouver, ISO, and other styles
38

Brownell, A., S. Lowson, and M. Brozovic. "Serum ferritin concentration in sickle cell crisis." Journal of Clinical Pathology 39, no. 3 (March 1, 1986): 253–55. http://dx.doi.org/10.1136/jcp.39.3.253.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Hellung-Larsen, P. "Parameters affecting the maximum cell concentration ofTetrahymena." Experientia 44, no. 1 (January 1988): 58–60. http://dx.doi.org/10.1007/bf01960245.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Dragoo, Jason L., and Malcolm R. DeBaun. "Stem Cell Yield after Bone Marrow Concentration." Orthopaedic Journal of Sports Medicine 5, no. 7_suppl6 (July 2017): 2325967117S0044. http://dx.doi.org/10.1177/2325967117s00445.

Full text
APA, Harvard, Vancouver, ISO, and other styles
41

van der Westhuizen, W. A., H. de Bruiyn, G. J. Beuckes, and J. C. Loock. "A separation cell for heavy-mineral concentration." Journal of Sedimentary Research 63, no. 4 (July 1, 1993): 765–66. http://dx.doi.org/10.1306/d4267bf9-2b26-11d7-8648000102c1865d.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Angomachalelis, Nestor J., Hercules S. Titopoulos, Manthos G. Tsoungas, and Agapios Gavrielides. "Red Cell Magnesium Concentration in Cor Pulmonale." Chest 103, no. 3 (March 1993): 751–55. http://dx.doi.org/10.1378/chest.103.3.751.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Ducommun, P., I. Bolzonella, M. Rhiel, P. Pugeaud, U. von Stockar, and I. W. Marison. "On-line determination of animal cell concentration." Biotechnology and Bioengineering 72, no. 5 (2001): 515–22. http://dx.doi.org/10.1002/1097-0290(20010305)72:5<515::aid-bit1015>3.0.co;2-q.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Kihara, Shigeharu, Daniel A. Hartzler, and Sergei Savikhin. "Oxygen Concentration Inside a Functioning Photosynthetic Cell." Biophysical Journal 106, no. 9 (May 2014): 1882–89. http://dx.doi.org/10.1016/j.bpj.2014.03.031.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Laila, Farida, Dedi Fardiaz, Nancy Dewi Yuliana, M. Rizal M. Damanik, and Fitriya Nur Annisa Dewi. "Methanol Extract of Coleus amboinicus (Lour) Exhibited Antiproliferative Activity and Induced Programmed Cell Death in Colon Cancer Cell WiDr." International Journal of Food Science 2020 (January 25, 2020): 1–12. http://dx.doi.org/10.1155/2020/9068326.

Full text
Abstract:
Coleus amboinicus(Lour) (CA) has been reported to possess many pharmacological activities. In this study, evaluation of cytotoxicity using brine shrimp lethality bioassay and MTT assay using WiDr cell lines was carried out. The expression of several genes responsible for programmed cell death of the methanol extract of CA was also investigated. The morphology of the cells undergoing apoptosis was detected using Hoechst staining assay. The gene expression of BAX, BCL2, P53, Caspase 1, 7, 8, and 9 of treated samples with different concentrations (10, 15, 25 & 50 µg/ml) were measured with RT PCR. The phytochemical profiles were investigated using LC MS. The results showed that the lethality concentration (LC50) of methanol extract using brine shrimp was 34.545 µg/ml and the extract exhibited good antiproliferative activity against cancer cells WiDr with IC50 value (8.598 ± 2.68 µg/ml) as compared to standard drug 5-fluorouracil (IC50 value 1.839 ± 0.03 µg/ml). There was apoptotic evidences from the morphology of treated cells. The expressions of BAX,P53, and Caspase 9 were upregulated in lower concentration of the extract (10 and 15 µg/ml) but downregulated in higher concentration (25 and 50 µg/ml). BCL2 as anti-apoptotic gene was downregulated in all concentrations. Caspase 1 and Caspase 7 were upregulated in high concentration (25 and 50 µg/ml), but downregulated in lower concentrations. These data provide a mode of cell death for the methanol extract of CA in low concentrations corresponding to apoptosis with intrinsic pathway. Many valuable compounds identified including caffeic acid, rosmarinic acid, malic acid, eicosapentanoic acid, benserazide, alpha-linolenic acid, betaine, Salvanolic B, 4-hydroxibenzoic acid and firulic acid have been previously reported as being active agents against many cancer cells. This study suggested that CA might become an effective ingredient for health-beneficial foods to prevent colon cancer.
APA, Harvard, Vancouver, ISO, and other styles
46

Kempe, Hermannus, Anne Schwabe, Frédéric Crémazy, Pernette J. Verschure, and Frank J. Bruggeman. "The volumes and transcript counts of single cells reveal concentration homeostasis and capture biological noise." Molecular Biology of the Cell 26, no. 4 (February 15, 2015): 797–804. http://dx.doi.org/10.1091/mbc.e14-08-1296.

Full text
Abstract:
Transcriptional stochasticity can be measured by counting the number of mRNA molecules per cell. Cell-to-cell variability is best captured in terms of concentration rather than molecule counts, because reaction rates depend on concentrations. We combined single-molecule mRNA counting with single-cell volume measurements to quantify the statistics of both transcript numbers and concentrations in human cells. We compared three cell clones that differ only in the genomic integration site of an identical constitutively expressed reporter gene. The transcript number per cell varied proportionally with cell volume in all three clones, indicating concentration homeostasis. We found that the cell-to-cell variability in the mRNA concentration is almost exclusively due to cell-to-cell variation in gene expression activity, whereas the cell-to-cell variation in mRNA number is larger, due to a significant contribution of cell volume variability. We concluded that the precise relationship between transcript number and cell volume sets the biological stochasticity of living cells. This study highlights the importance of the quantitative measurement of transcript concentrations in studies of cell-to-cell variability in biology.
APA, Harvard, Vancouver, ISO, and other styles
47

Naundorf, Gerardo, and Nicholas G. Aumen. "Ammonia-induced cell envelope injury in Escherichia coli and Enterobacter aerogenes." Canadian Journal of Microbiology 36, no. 8 (August 1, 1990): 525–29. http://dx.doi.org/10.1139/m90-092.

Full text
Abstract:
Ammonia-induced cell envelope injury was examined in pure cultures of Escherichia coli and Enterobacter aerogenes. Cell injury, as determined by the ratio of colony-forming units on m-T7 agar to colony-forming units on m-Endo agar, increased with exposure to increasing concentrations of ammonia. Cell envelopes appeared to be the site of injury as indicated by increasing susceptibility to lysozyme with increasing ammonia concentration. Cells exposed to ammonia also exhibited more cellular leakage than control cells. Leakage from cells exposed to ammonia included proteins, and all leaked substances increased in concentration as ammonia concentrations increased. The concentration of 2-keto-3-deoxyoctonate (KDO) in the outer membrane of E. coli increased with ammonia exposure, while KDO concentration in the outer membrane of E. aerogenes decreased. The results suggest that exposure of E. coli cells to high concentrations of ammonia disrupts the outer membrane and lipopolysaccharide-associated proteins, while E. aerogenes cells are affected through the disruption of bonds between KDO and the outer membrane. Key words: injury, coliform, ammonia, cell envelope.
APA, Harvard, Vancouver, ISO, and other styles
48

Ściskalska, Milena, Grzegorz Marek, Zygmunt Grzebieniak, and Halina Milnerowicz. "Resistin as a Prooxidant Factor and Predictor of Endothelium Damage in Patients with Mild Acute Pancreatitis Exposed to Tobacco Smoke Xenobiotics." Mediators of Inflammation 2017 (2017): 1–10. http://dx.doi.org/10.1155/2017/3039765.

Full text
Abstract:
Objectives. The study was aimed to assess the influence of tobacco smoke exposure on the intensity of inflammation measured by IL-6, α1-antitripsin (AAT) and α1-acid glycoprotein (AGP) concentrations, and Cd level and oxidative stress intensity measured by advanced oxidation protein product (AOPP) concentration in the blood of healthy subjects and AP patients during hospitalization. Endothelin-1 (ET-1) and resistin concentrations, markers of endothelium injury, were determined. Results. An increased IL-6 concentration in healthy smokers compared to nonsmokers and AP patients compared to controls was shown. An increased AAT and AGP concentrations during hospitalization of AP patients were noted, in both smokers (AAT, AGP) and nonsmokers (AAT). In comparison to control groups, in AP patients, a 2-fold increased resistin concentration correlating with ET-1 concentration and decreased albumin concentration accompanied by increased AOPP concentration were demonstrated. AOPP concentration was higher in smokers with AP compared to nonsmokers and gradually enhanced during their hospitalization. Conclusions. Tobacco smoke exposure can have a proinflammatory effect in both healthy subjects and AP patients. Increased resistin concentration in AP patients negatively correlating with albumin concentration has prooxidative effect on this protein resulting in enhanced AOPP level. Increased resistin concentration can intensify AAT and AGP production during AP.
APA, Harvard, Vancouver, ISO, and other styles
49

Ren, Kai, Yong X. Gan, Efstratios Nikolaidis, Sharaf Al Sofyani, and Lihua Zhang. "Electrolyte Concentration Effect of a Photoelectrochemical Cell Consisting of TiO2 Nanotube Anode." ISRN Materials Science 2013 (March 20, 2013): 1–7. http://dx.doi.org/10.1155/2013/682516.

Full text
Abstract:
The photoelectrochemical responses of a TiO2 nanotube anode in ethylene glycol (EG), glycerol, ammonia, ethanol, urea, and Na2S electrolytes with different concentrations were investigated. The TiO2 nanotube anode was highly efficient in photoelectrocatalysis in these solutions under UV light illumination. The photocurrent density is obviously affected by the concentration change. Na2S generated the highest photocurrent density at 0, 1, and 2 V bias voltages, but its concentration does not significantly affect the photocurrent density. Urea shows high open circuit voltage at proper concentration and low photocurrent at different concentrations. Externally applied bias voltage is also an important factor that changes the photoelectrochemical reaction process. In view of the open circuit voltage, EG, ammonia, and ethanol fuel cells show the trend that the open circuit voltage (OCV) increases with the increase of the concentration of the solutions. Glycerol has the highest OCV compared with others, and it deceases with the increase in the concentration because of the high viscosity. The OCV of the urea and Na2S solutions did not show obvious concentration effect.
APA, Harvard, Vancouver, ISO, and other styles
50

Zhao, Yu, Xiao Bin Wang, Peng Li, and Yan Ping Sun. "The Influence of Mediator Concentration on the Performance of Microbial Fuel Cell." Advanced Materials Research 512-515 (May 2012): 1520–24. http://dx.doi.org/10.4028/www.scientific.net/amr.512-515.1520.

Full text
Abstract:
Electrochemical impedance spectroscopy (EIS), cyclic voltammetry (CV), power density and anode potential are used to characterize the mediator microbial fuel cell at different methylene blue (MB) concentrations. At lower MB concentration between 9.98×10-3 mmol/L and 1.66×10-1 mmol/L, the increased power density is enabled by using high mediator concentrations. Higher peak power density of 159.6 mw/m2 is observed compared with the peak power density of 36.0 mw/m2. But MB at too high concentration is disadvantageous to the perform of MFC. At the MB concentration of 2.50×10-1 mmol/L, the peak power output is just 128.4 mw/m2, which is lower than 159.6 mw/m2 at MB concentration of 1.66×10-1 mmol/L.
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography