首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 140 毫秒
1.

Background

Previous studies of compatibility of calcium chloride (CaCl2) and phosphates have not included particle counts in the range specified by the United States Pharmacopeia. Micro-flow imaging techniques have been shown to be comparable to light obscuration when determining particle count and size in pharmaceutical solutions.

Objective

The purpose of this study was to do compatibility testing for parenteral nutrition (PN) solutions containing CaCl2 using dynamic light scattering and micro-flow imaging techniques.

Methods

Solutions containing TrophAmine (Braun Medical Inc, Irvine, CA), CaCl2, and sodium phosphate (NaPhos) were compounded with and without cysteine. All solutions contained standard additives to neonatal PN solutions including dextrose, trace metals, and electrolytes. Control solutions contained no calcium or phosphate. Solutions were analyzed for particle size and particle count. Means of Z-average particle size and particle counts of controls were determined. Study solutions were compared to controls and United States Pharmacopeia (USP) Chapter 788 guidelines. The maximum amount of Phos that was compatible in solutions that contained at least 10 mmol/L of Ca in 2.5% amino acids (AA) was determined. Compatibility of these solutions was verified by performing analyses of 5 repeats of these solutions. Microscopic analyses of the repeats were also performed.

Results

Amounts of CaCl2 and NaPhos that were compatible in solutions containing 1.5%, 2%, 2.5%, and 3% AA were determined. The maximum amount of NaPhos that could be added to TrophAmine solutions of > = 2.5% AA containing at least 10 mmol/L of CaCl2 was 7.5 mmol/L. Adding 50 mg/dL of cysteine increased the amount of NaPhos that could be added to solutions containing 10 mmol/L of CaCl2 to 10 mmol/L.

Conclusion

Calcium chloride can be added to neonatal PN solutions containing NaPhos in concentrations that can potentially provide an intravenous intake of adequate amounts of calcium and phosphorus.  相似文献   

2.
We studied the extraction and analysis of integral membrane proteins possessing hydrophobic and hydrophilic domains and found that a nonionic detergent called MEGA-10, used in lysis buffers, had a superior extraction effect compared to most conventional detergents. A sodium dodecyl sulfate (SDS) concentration of >0.4% (w/v) in the sample buffer was crucial for those proteins to be clearly analyzed by electrophoresis and Western blotting. Furthermore, MEGA-10 had the tendency to maximally extract proteins around its critical micelle concentration (CMC) of 0.24% (w/v). These solutions can greatly assist functional investigations of membrane proteins in the proteomics era.  相似文献   

3.
In this study, the effect of concentration (1/2 CMC, at CMC and 2 x CMC) of surfactants, cetyl trimethyl ammonium bromide (cationic), sodium dodecyl sulfate (anionic), and tween ‘20’ (non-ionic) on the movement of carbofuran, chlorpyrifos and en-dosulfan in soils was evaluated by using a soil thin-layer chromatographic technique. The movement of pesticides was detected by spray reagents and expressed in terms of Rf values. The penetrability K was found to increase by decreasing the plate angle and followed the order as: sandy loam > loam > silt loam soils. The penetrability K also decreases in surfactant-free and surfactant-amended soils when developed in distilled water and aqueous surfactant solutions of different CMCs, respectively. The higher movement of pesticides was observed in sandy loam soil followed by loam and silt loam soils. On the basis of Rf values, the movement of pesticides follows the order as: carbofuran > chlorpyrifos > endosulfan, both in surfactant-amended and surfactant-free soils when developed in distilled water and aqueous surfactant solutions of different CMCs. The movement is directly proportional to the aqueous solubilities, polarities, and carbon numbers and inversely related to the molecular weights of pesticides. A significant increase or decrease of pesticides movement in soils was discussed on the basis of adsorption of pesticides on soils, chemical nature of the surfactants, and its concentrations in terms of critical micelle concentrations (CMCs) in soils and eluents. Results obtained may provide insights pertaining to the use of surfactants for solving soil pollution problems posed by pesticides.  相似文献   

4.
The influence of the temperature on the adsorption of monomeric and micellar solutions of the anionic surfactant sodium dodecyl sulfate (SDS) on phosphatidylcholine (PC) liposomes was investigated using the fluorescent probe 2-(p-toluidinyl)-naphthalene-6-sodium sulfonate (TNS). The number of adsorbed molecules was quantified by measuring changes in the electrostatic potential (Psi(o)) of the liposomes/probe during an incubation with SDS at varying temperatures. At low surfactant concentrations (from 0.05 to 0.25 mM), the increase in temperature reduced the number of surfactant molecules incorporated per vesicle regardless of the incubation time, whereas at high surfactant concentrations (from 0.50 to 1.0 mM) the incubation time has an opposite effect on this process. Thus, after 10s, the surfactant adsorption decreased with temperature, yet it increased progressively with time. The adsorption was linear with temperature below critical micellar concentration (CMC) of SDS and this linear tendency did not change above CMC. This suggests an adsorption of SDS monomers regardless of the surfactant concentration.  相似文献   

5.
Summary The extraction of penicillin acylase by reverse micellar solutions of a surfactant was studied. A 50 mM solution of dioctyl sodium sulphosuccinate in isooctane extracted 46% of the enzyme activity in a crude periplasmic extract of induced cells of E. coli ATCC 9637. The increase in the specific activity of the final enzyme preparation, after stripping of the organic phase at pH 7.5, in the presence of 1 M KCl, was 8 - fold.Abbreviations PA penicillin acylase (penicillin amidohydrolase EC 3.5.1.11) - AOT Aerosol OT (dioctyl sodium sulphosuccinate) - NIPAB 6-nitro-3-(phenylacetamido)-benzoic acid - NABA 6-nitro-3-aminobenzoic acid - BSA bovine serum albumin - SDS sodium dodecylsulphate  相似文献   

6.
A study has been made of the effect of sodium dodecylsufate (SDS) addition on the oxidation of ethanol catalyzed by yeast alcohol dehydrogenase. Experiments were performed at pH = 8.1 and SDS concentrations employed were below and above the surfactant critical micelle concentration (CMC). The double reciprocal plots obtained in the absence and in the presence of the surfactant were compatible with a sequential bi-bi ordered mechanism. In the presence of the surfactant the initial reaction rates were consistently lower than in pure buffer at all the surfactant concentrations considered (0.5-50 mM). This effect is mainly due to an increase in the dissociation constant of beta-NAD(+) which reaches its maximum value (7,100 +/- 1,700 microM) at the CMC. Above the CMC the effect of the surfactant is mainly due to an increase in the Michaels constants of the alcohol, with values of 41 +/- 1 mM for 15 mM SDS and 50 +/- 1 mM for 50 mM SDS. The catalytic rate constant was found to be practically independent of the presence of the surfactant in the range of concentrations considered (up to 50 mM).  相似文献   

7.
This investigation examined the effects of common aqueous biocides and disinfectant foams derived from them on Pseudomonas aeruginosa biofilms. Biofilms were grown on stainless steel coupons under standardised conditions in a reactor supplemented with low concentrations of organic matter to simulate conditions prevalent in industrial systems. Five-day-old biofilms formed under ambient conditions with continuous agitation demonstrated a low coefficient of variation (5.809%) amongst viable biofilm bacteria from independent trials. Scanning electron microscopy revealed biofilms on coupons with viable biofilm bacteria observed by confocal microscopy. An aqueous solution of a common foaming agent amine oxide (AO) produced negligible effects on bacterial viability in biofilms (p?>?0.05). However, significant biofilm inactivation was noted with aqueous solutions of common biocides (peracetic acid, sodium hypochlorite, sodium ethylenediaminetetraacetic acid) with or without AO (p?<?0.05). Aereation of a mixture of AO with each of these common biocides resulted in significant reductions in the viability of biofilm bacteria (p?<?0.05). In contrast, limited effects were noted by foam devoid of biocides. A relationship between microbial inactivation and the concentration of biocide in foam (ranging from 0.1?–?0.5%) and exposure period were noted (p?<?0.05). Although, lower numbers of viable biofilm bacteria were recovered after treatment with the disinfectant foam than by the cognate aqueous biocide, significant differences between these treatments were not evident (p?>?0.05). In summary, the studies revealed significant biofilm inactivation by biocidal foam prepared with common biocides. Validation of foam disinfectants in controlled trials at manufacturing sites may facilitate developments for clean in place applications. Advantages of foam disinfectants include reductions in the volumes of biocides for industrial disinfection and in their disposal after use.  相似文献   

8.
A membrane-covered polarographic oxygen electrode was used to measure oxygen diffusion coefficients in aqueous polyelectrolyte solutions of xanthan gum, sodium alginate, and sodium carboxymethylcellulose (CMC). In sodium alginate solutions, dilute xanthan solutions, and solutions containing more than 0.3 wt % CMC, oxygen diffusion coefficients decrease with increasing polymer concentrations. Interestingly, in dilute CMC solutions and concentrate xanthan solutions containing more than 0.5 wt % xanthan gum, oxygen diffusion coefficients increase with increasing polymer concentrations, and values exceeding that in pure water are generally observed.  相似文献   

9.
The effect of different concentrations of Na2SeO3 on human pulmonary adenocarcinoma cells and human embryonic lung diploid cells in vitro was investigated. For human pulmonary adenocarcinoma cells, mitotic index and cell count decreased with increasing selenium concentrations. At 1 μg/mL sodium selenite, mitotic activity and growth of human lung cancer cells were partially inhibited, and the progression of human lung cancer cell cycle was partially arrested. When human embryonic lung diploid cells were treated with 1 μg/mL sodium selenite for five continuous days, cell counts of the treated group were closely parallel to those of the control group. After treating human embryonic lung diploid cells with 1–5μg/mL sodium selenite for 1–3 d, the mitotic index (MI), labeled index (LI), and average silver grain (SG) number per 20 labeled nuclei were the same as those of the control. In mixed cultures of human embryonic lung diploid cells and human pulmonary adenocarcinoma cells, treated with 3 and 5 μg/mL sodium selenite for 24 h, the lung diploid cells showed a normal fusiform morphology, whereas the lung cancer cells showed heavily vacuolated cytoplasms and distorted nuclei.  相似文献   

10.
Naturally contaminated lots of Canadian barley containing either 18.4 or 4.3 μg/g deoxynivalenol (DON) were heated at 80 °C, with small amounts of water or 1 M sodium carbonate solution to study the rate of DON reduction. Samples were heated in sealed polypropylene containers for periods of up to 8 days. In the 18.4 μg/g DON barley, rapid reductions were observed: with no solutions added, DON declined to 14.7 μg/g after 1 day, and to 4.9 μg/g after 8 days solely due to heat; with water at 10 mL/100 g barley, DON levels reached 3.7 μg/g after 8 days; with 1 M sodium carbonate solution added at 10 mL/100 g barley, DON declined to 4.7 μg/g after 1 day, and to 0.4 μg/g after 8 days; with 20 mL/100 g barley, DON declined to 1.4 μg/g after 1 day and to near-zero levels after 8 days. In the 4.3 μg/g DON barley, more gradual reductions were evident: with no solutions added, DON declined to 2.9 μg/g after 8 days solely due to heat; with water at 10 mL/100 g barley, DON levels reached 2.3 μg/g after 8 days; with 1 M sodium carbonate solution added at 10 mL/100 g barley, DON declined to 2.7 μg/g after 1 day, and to near-zero levels after 8 days; with 20 mL/100 g barley, DON declined to 1.4 μg/g after 1 day and to near-zero levels after 3, 5 and 8 days.  相似文献   

11.
Extracellular vesicles (EVs) have emerged as a promising biomarker platform for glioblastoma patients. However, the optimal method for quantitative assessment of EVs in clinical bio-fluid remains a point of contention. Multiple high-resolution platforms for quantitative EV analysis have emerged, including methods grounded in diffraction measurement of Brownian motion (NTA), tunable resistive pulse sensing (TRPS), vesicle flow cytometry (VFC), and transmission electron microscopy (TEM). Here we compared quantitative EV assessment using cerebrospinal fluids derived from glioblastoma patients using these methods. For EVs <150 nm in diameter, NTA detected more EVs than TRPS in three of the four samples tested. VFC particle counts are consistently 2–3 fold lower than NTA and TRPS, suggesting contribution of protein aggregates or other non-lipid particles to particle count by these platforms. While TEM yield meaningful data in terms of the morphology, its particle count are consistently two orders of magnitude lower relative to counts generated by NTA and TRPS. For larger particles (>150 nm in diameter), NTA consistently detected lower number of EVs relative to TRPS. These results unveil the strength and pitfalls of each quantitative method alone for assessing EVs derived from clinical cerebrospinal fluids and suggest that thoughtful synthesis of multi-platform quantitation will be required to guide meaningful clinical investigations.  相似文献   

12.
Forming a salt is a common way to increase the solubility of a poorly soluble compound. However, the solubility enhancement gained by salt formation may be lost due to solution-mediated phase transformation (SMPT) during dissolution. The SMPT of a salt can occur due to a supersaturated solution near the dissolving surface caused by pH or other solution conditions. In addition to changes in pH, surfactants are also known to affect SMPT. In this study, SMPT of a highly soluble salt, haloperidol mesylate, at pH 7 in the presence of a commonly used surfactant, sodium lauryl sulfate (SLS), was investigated. Dissolution experiments were performed using a flow-through dissolution apparatus with solutions containing various concentrations of SLS. Compacts of haloperidol mesylate were observed during dissolution in the flow-through apparatus using a stereomicroscope. Raman microscopy was used to characterize solids. The dissolution of haloperidol mesylate was significantly influenced by the addition of sodium lauryl sulfate. In conditions where SMPT was expected, the addition of SLS at low concentrations (0.1–0.2 mM) reduced the dissolution of haloperidol mesylate. In solutions containing concentrations of SLS above the critical micelle concentration (CMC) (10–15 mM), the dissolution of haloperidol mesylate increased compared to below the CMC. The solids recovered from solubility experiments of haloperidol mesylate indicated that haloperidol free base precipitated at all concentrations of SLS. Above 5 mM of SLS, Raman microscopy suggested a new form, perhaps the estolate salt. The addition of surfactant in solids that undergo solution-mediated phase transformation can add complexity to the dissolution profiles and conversion.  相似文献   

13.
Bioremediation efforts often rely on the application of surfactants to enhance hydrocarbon bioavailability. However, synthetic surfactants can sometimes be toxic to degrading microorganisms, thus reducing the clearance rate of the pollutant. Therefore, surfactant-resistant bacteria can be an important tool for bioremediation efforts of hydrophobic pollutants, circumventing the toxicity of synthetic surfactants that often delay microbial bioremediation of these contaminants. In this study, we screened a natural surfactant-rich compartment, the estuarine surface microlayer (SML), for cultivable surfactant-resistant bacteria using selective cultures of sodium dodecyl sulfate (SDS) and cetyl trimethylammonium bromide (CTAB). Resistance to surfactants was evaluated by colony counts in solid media amended with critical micelle concentrations (CMC) of either surfactants, in comparison with non-amended controls. Selective cultures for surfactant-resistant bacteria were prepared in mineral medium also containing CMC concentrations of either CTAB or SDS. The surfactantresistant isolates obtained were tested by PCR for the Pseudomonas genus marker gacA gene and for the naphthalene-dioxygenase-encoding gene ndo. Isolates were also screened for biosurfactant production by the atomized oil assay. A high proportion of culturable bacterioneuston was tolerant to CMC concentrations of SDS or CTAB. The gacA-targeted PCR revealed that 64% of the isolates were Pseudomonads. Biosurfactant production in solid medium was detected in 9.4% of tested isolates, all affiliated with genus Pseudomonas. This study shows that the SML is a potential source of surfactant-resistant and biosurfactant-producing bacteria in which Pseudomonads emerge as a relevant group.  相似文献   

14.
This study describes the interaction between sodium dodecyl sulfate (SDS) and membrane proteins reconstituted into large unilamellar lipid vesicles and detergent micelles studied by circular dichroism (CD) and polarity sensitive probe labeling. Specifically, we carried out a comparative study of two aquaporins with high structural homology SoPIP2;1 and AqpZ using identical reconstitution conditions. Our CD results indicate that SDS, when added to membrane-reconstituted aquaporins in concentrations below the SDS critical micelle concentration (CMC, ~8mM), causes helical rearrangements of both aquaporins. However, we do not find compelling evidence for unfolding. In contrast when SDS is added to detergent stabilized aquaporins, SoPIP2;1 partly unfolds, while AqpZ secondary structure is unaffected. Using a fluorescent polarity sensitive probe (Badan) we show that SDS action on membrane reconstituted SoPIP2;1 as well as AqpZ is associated with initial increased hydrophobic interactions in protein transmembrane (TM) spanning regions up to a concentration of 0.1× CMC. At higher SDS concentrations TM hydrophobic interactions, as reported by Badan, decrease and reach a plateau from SDS CMC up to 12.5× CMC. Combined, our results show that SDS does not unfold neither SoPIP2;1 nor AqpZ during transition from a membrane reconstituted form to a detergent stabilized state albeit the native folds are changed.  相似文献   

15.
We describe a simple and rapid method for determining the critical micelle concentration (CMC) of surfactants from fiber-optic measurements of refractive index. The refractive index of an aqueous surfactant solution was monitored as the surfactant concentration was increased using an automated dispensing system. On reaching the surfactant’s CMC value, an abrupt change was observed in the rate of increase of the refractive index with increasing concentration. The measurement system provides rapid semiautomatic data collection and analysis, increasing the precision, sensitivity, and range of applicability of the technique while substantially decreasing the amount of manual intervention required. Measurements of CMC for sodium dodecyl sulfate (8.10 mM), cetyltrimethylammonium chloride (1.58 mM), and Triton X-100 (0.21 mM) were in excellent agreement with values previously reported in the literature. The method is applicable to cationic, anionic, and nonionic surfactants, and it offers a facile, in situ, and sensitive means of detecting micelle formation over a broad range of CMC values larger than 10−1 mM.  相似文献   

16.
Voltage clamp studies were performed on the dorsal integument of Hirudo medicinalis. Under apical calcium-free conditions an inward-directed component of transepithelial current was activated by changes of transepithelial voltage. Depolarization caused up to 50% increase of the transepithelial sodium current. Hyperpolarization had no comparable effects. With calcium (1.8 mM) or amiloride (100 μM) in the apical solution and in sodium-free solutions the inward-directed current failed to increase after depolarization. Activation also occurred under chloride-free conditions. Permeabilization of the apical membrane by nystatin (5 μM) increased the current activation significantly. After nystatin, calcium as well as amiloride lost their inhibitory effects. This indicates a basolateral localization of the voltage-dependent conductance. Vesicle insertion or cytoskeletal structures are probably not involved in regulation, as seen by the lack of effects of brefeldin A and the cytochalasins B and D. However, serosal hyposmolar solutions (170 mosmol · l−1) caused a reinforced activation of the current. Our results indicate a voltage-dependent conductance in a tight sodium-absorbing epithelium. Accepted: 22 January 1998  相似文献   

17.
A method has been developed to monitor the cell growth of aggregated human embryonic kidney 293 (HEK293) suspension cultures by measuring cumulative particle volume and the particle size distribution. This method employs a particle size analyzer that determines the size of individual particles by detecting their light obscuration (blockage) or scattering. Cell counts derived from the cumulative volume of the cell particle correlate well with manual cell counts from a hemacytometer at different stages of growth. This correlation was further confirmed by quantifying total cellular protein of the samples. Simultaneously, the aggregation state of the samples can also be monitored and mathematically described. Results from this study demonstrate that this simple and reproducible method allows the direct measurement of cumulative cell volume and the degree of cell aggregation, as well as an indirect assessment of cell counts.  相似文献   

18.
The preseparator of an Andersen impactor with different coating treatments for a range of particle-size distributions was evaluated. Limited theoretical simulations constrained by simplifying assumptions of the airflow fields in the preseparator and upper stages of an 8-stage Andersen impactor were used to reveal low-velocity and high-pressure regions for potential deposition. These regions were then sampled in subsequent particle deposition experiments. Disodium fluorescein aerosols were sampled with different coating treatments of the preseparator floor. Particles collected at impactor stages determined particle size distributions. Stage deposition was compared between different preseparator treatments (buffer and silicon oil). Collection efficiency in the preseparator followed the pattern buffer >silicon oil >untreated. Statistical differences (P>0.05) were noted in collection efficiency of large particles (45 μm-75 μm) in the preseparator. The mass median aerodynamic diameters and geometric standard deviations showed some statistical differences when different preseparator treatments for large particles were used; therefore, preseparator coating was shown to influence performance and thereby estimates of particle size by intertial impaction.  相似文献   

19.
The recovery of serum proteins from reversed-phase and internal-surface reversed-phase (ISRP) silica supports following direct serum injection was investigated using an eluent containing a micellar solution of sodium dodecyl sulphate (SDS). The results indicated that the recoveries of serum proteins were 98–103% for both supports. On the basis of the above findings, the separation and recovery of hydrophilic drugs (cephalosporins and salicylic acid) from human serum were investigated using acidic eluents including micellar solutions of SDS. They were completely separated from the components of serum, and the recoveries were 94–98% despite protein binding. Although the recommended eluent pH range is 6.0–7.5 for the ISRP support, eluents of pH 2–8 can be used with the micellar chromatographic system.  相似文献   

20.
胶体微晶纤维素(avicel)是一种由微晶纤维素(microcrystalline cellulose, MCC)和羧甲基纤维素(carboxymethyl cellulose,CMC)制成的混合物,可用于病毒蚀斑测定。常用的avicel由FMC公司生产,其MCC和CMC比例相对固定,无法很好地适应所有类型病毒的蚀斑测定实验。本研究通过对比不同的MCC和CMC配制比例对avicel在病毒蚀斑测定作用的影响,建立了一种操作简便、实用性好和稳定性好的改良avicel病毒蚀斑测定法。为了配制不同浓度MCC和CMC的混合物,本研究制备出12种2×avicel覆盖层,测定其总体黏度及底层黏度,评估其与传统覆盖层相比,使用时的操作难易程度。进一步将12种2×avicel覆盖层制备成avicel-DMEM营养覆盖层,测定96孔板中猪流行性腹泻病毒滴度,比较12种avicel覆盖层及传统覆盖层蚀斑大小、清晰度、稳定性及滴度准确性等的差异,筛选出最佳测定方法。结果显示,12种2×avicel覆盖层中,除4.8%MCC+1.4%CMC和4.8%MCC+1.0%CMC外,其余2×avicel覆盖层在实际使用中均比2×CMC覆盖层更容易吸取和配制营养覆盖层。最后,利用avicel病毒蚀斑测定法测定96孔板中猪流行性腹泻病毒滴度,结果显示CMC浓度越高蚀斑越小,其中终浓度为0.6%MCC+0.7%CMC的avicel覆盖层测定蚀斑染色最清晰,准确度与传统覆盖层相似,但操作较传统覆盖层更简便。综上所述,本研究建立了一种操作简便、实用性好和稳定性好的改良avicel病毒蚀斑测定法,为病毒的病原学、抗病毒药物及疫苗等相关研究的展开提供了良好的实验基础。  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号