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 31 毫秒
1.
2.
c‐MYC overexpression is frequently observed in various cancers including colon cancer and regulates many biological activities such as aberrant cell proliferation, apoptosis, genomic instability, immortalization and drug resistance. However, the mechanism by which c‐MYC confers drug resistance remains to be fully elucidated. In this study, we found that the c‐MYC expression level in primary colorectal cancer tissues correlated with the recurrence rate following 5‐fluorouracil (5‐FU)‐based adjuvant chemotherapy. Supporting this finding, overexpression of exogenous c‐MYC increased the survival rate following 5‐FU treatment in human colon cancer cells, and knockdown of endogenous c‐MYC decreased it. Furthermore, c‐MYC knockdown decreased the expression level of ABCB5, which is involved in 5‐FU resistance. Using a chromatin immunoprecipitation assay, we found that c‐MYC bound to the ABCB5 promoter region. c‐MYC inhibitor (10058‐F4) treatment inhibited c‐MYC binding to the ABCB5 promoter, leading to a decrease in ABCB5 expression level. ABCB5 knockdown decreased the survival rate following 5‐FU treatment as expected, and the ABCB5 expression level was increased in 5‐FU‐resistant human colon cancer cells. Finally, using a human colon cancer xenograft murine model, we found that the combined 5‐FU and 10058‐F4 treatment significantly decreased tumorigenicity in nude mice compared with 5‐FU or 10058‐F4 treatment alone. 10058‐F4 treatment decreased the ABCB5 expression level in the presence or absence of 5‐FU. In contrast, 5‐FU treatment alone increased the ABCB5 expression level. Taken together, these results suggest that c‐MYC confers resistance to 5‐FU through regulating ABCB5 expression in human colon cancer cells.  相似文献   

3.
Gynostemma pentaphyllum saponins (GpS) have been shown to have anti‐cancer activity. However, the underlying mechanisms remain unclear. In this study, we used the ApcMin/+ colorectal cancer (CRC) mouse model to investigate the anti‐cancer effect of GpS and we demonstrated that GpS treatment could significantly reduce the number and size of intestinal polyps in ApcMin/+ mice. In order to identify the potential targets and mechanisms involved, a comparative proteomics analysis was performed and 40 differentially expressed proteins after GpS treatment were identified. Bioinformatics analyses suggested a majority of these proteins were involved in processes related to cellular redox homeostasis, and predicted Raf‐1 as a potential target of GpS. The upregulation of two proteins known to be involved in redox homeostasis, peroxiredoxin‐1 (Prdx1) and peroxiredoxin‐2 (Prdx2), and the downregulation of Raf‐1 were validated using Western blot analysis. After further investigation of the associated signaling networks, we postulated that the anti‐cancer effect of GpS was mediated through the upregulation of Prdx1 and Prdx2, suppression of Ras, RAF/MEK/ERK/STAT, PI3K/AKT/mTOR signaling and modulation of JNK/p38 MAPK signaling. We also examined the potential combinatorial effect of GpS with the chemotherapeutic 5‐fluorouracil (5‐FU) and found that GpS could enhance the anti‐cancer efficacy of 5‐FU, further suppressing the number of polyps in ApcMin/+ mice. Our findings highlight the potential of GpS as an anti‐cancer agent, the potential mechanisms of its anti‐cancer activities, and its effect as an adjuvant of 5‐FU in the chemotherapy of CRC.  相似文献   

4.
The membrane potential in highly coupled potato ( Solanum tuberosum L.) mitochondria, as measured by changes in safranine absorbance, was significantly increased by addition of bovine serum albumin. Purification of potato mitochondria on Percoll, in removing 50% of free unsaturated fatty acids, decreased the BSA-de-pendent membrane potential. The effect of added linoleic acid and of the natural accumulation of fatty acids during aging was studied. The response of membrane potential to addition of bovine serum albumin appeared to be directly correlated to the amount of free unsaturated fatty acids. Aging in vitro, in releasing free fatty acids, decreased respiratory control and ADP:O ratios and collapsed the membrane potential. During 2–3 h of incubation, addition of BSA completely restored membrane potential and oxidative phosphorylation.
It is concluded that both in fresh and in aged potato mitochondria the effect of bovine serum albumin on oxidative phosphorylation can be ascribed to an effect on membrane permeability to ions. BSA, in binding free unsaturated fatty acids, restored maximal membrane potential. The bovine serum albumin-dependent membrane potential appears to be a sensitive criterion of the functional integrity of the inner mitochondrial membrane.  相似文献   

5.
Interaction of 3‐styrylindoles 1–8 viz. 3‐(2‐phenylethenyl‐E)‐NH‐indole (1), 3‐[2‐(4‐nitrophenyl)ethenyl‐E]‐NH‐indole (2), 5‐bromo‐3‐[2‐(4‐nitrophenyl)ethenyl‐E]‐NH‐indole (3), 5‐methoxy‐3‐[2‐(4‐nitrophenyl)ethenyl‐E]‐NH‐indole (4), 3‐[2‐(4‐cyanophenyl)ethenyl‐E]‐NH‐indole (5), 3‐[2‐(4‐cyanophenyl)ethenyl‐E]‐N‐ethylindole (6), 5‐bromo‐3‐[2‐(4‐chlorophenyl)ethenyl‐E]‐NH‐indole (7) and 5‐methoxy‐3‐[2‐(4‐chlorophenyl)ethenyl‐E]‐NH‐indole (8) with bovine serum albumin (BSA) was examined by UV–vis and steady‐state fluorescence spectroscopy. The fluorescence intensity of 1–8 increases with the increasing BSA concentration. Upon binding with BSA, while 1 and 5–8 show a blue shift in their λf max, 2–4 do not exhibit such behavior. Compounds 1–8 also quench the 345 nm fluorescence of BSA in phosphate buffer (λex, 280 nm). These compounds intercalate in the hydrophobic regions of BSA, as evidenced by the determination of BSA binding site micropolarity using compounds 2–8. As evidenced by the estimation of energy transfer efficiency and distance between the donor (BSA‐Trp‐212) and the acceptor (3‐styrylindoles), the halo‐substituted compounds 3 and 7 interact with BSA more effectively than the other 3‐strylindoles. These compounds have potential for use as neutral and hydrophobic fluorescence probes for examining the microenvironments in proteins, polymers, micelles, etc. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

6.
The indole derivative 2-(5-methoxy-2-methyl-1H-indol-3-yl)-N'-[(E)-(3-nitrophenyl) methylidene]acetohydrazide (IND) was synthesized for its therapeutic potential to inhibit cyclooxygenase (COX)-II. Binding if IND to bovine serum albumin (BSA) was investigated was because most drugs bind to serum albumin in-vivo. Fluorescence, UV–vis spectrophotometry and molecular modeling methodologies were employed for studying the interaction mechanism. The intrinsic fluorescence of BSA was quenched by BSA and the quenching mechanism involved was static quenching. The binding constants between IND and BSA at the three studied temperatures (298, 301 and 306 K) were 1.09 × 105, 4.36 × 104 and 1.23 × 104 L mol−1 respectively. The most likely site for binding IND to BSA was Site I (subdomain IIA). The analysis of thermodynamic parameter revealed the involvement of hydrogen bonding and van der Waals forces in the IND-BSA interaction. Synchronous fluorescence spectroscopic (SFS) and UV–vis spectrophotometric studies suggested conformational change in BSA molecule post interaction to IND. Molecular docking and the experimental results corroborated one another. The study can prove as an insight for future IND drug development.  相似文献   

7.
The interaction between the photosensitive antitumour drug, 2(3),9(10),16(17),23(24)‐tetra‐(((2‐aminoethylamino)methyl)phenoxy)phthalocyaninato‐zinc(II) (ZnPc) and bovine serum albumin (BSA) has been investigated using various spectroscopic methods. This work may provide some useful information for understanding the interaction mechanism of anticancer drug–albumin binding and gain insight into the biological activity and metabolism of the drug in blood. Based on analysis of the fluorescence spectra, ZnPc could quench the intrinsic fluorescence of BSA and the quenching mechanism was static by forming a ground state complex. Meanwhile, the Stern–Volmer quenching constant (KSV), binding constant (Kb), number of binding sites (n) and thermodynamic parameters were obtained. Results showed that the interaction of ZnPc with BSA occurred spontaneously via hydrogen bond and van der Waal's force. According to Foster's non‐radioactive energy transfer theory, the energy transfer from BSA to ZnPc occurred with high possibility. Synchronous fluorescence and circular dichroism (CD) spectra also demonstrated that ZnPc induced the secondary structure of and conformation changes in BSA, especially α helix. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

8.
Several spectroscopic approaches namely fluorescence, time‐resolved fluorescence, UV‐visible, and Fourier transform infra‐red (FT‐IR) spectroscopy were employed to examine the interaction between ethane‐1,2‐diyl bis(N,N‐dimethyl‐N‐hexadecylammoniumacetoxy)dichloride (16‐E2‐16) and bovine serum albumin (BSA). Fluorescence studies revealed that 16‐E2‐16 quenched the BSA fluorescence through a static quenching mechanism, which was further confirmed by UV–visible and time‐resolved fluorescence spectroscopy. In addition, the binding constant and the number of binding sites were also calculated. The thermodynamic parameters at different temperatures (298 K, 303 K, 308 K and 313 K) indicated that 16‐E2‐16 binding to BSA is entropy driven and that the major driving forces are electrostatic interactions. Decrease of the α‐helix from 53.90 to 46.20% with an increase in random structure from 22.56 to 30.61% were also observed by FT‐IR. Furthermore, the molecular docking results revealed that 16‐E2‐16 binds predominantly by electrostatic and hydrophobic forces to some residues in the BSA sub‐domains IIA and IIIA. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

9.
Anthocyanin is one of the flavonoid phytopigments that shows strong antioxidant activity. The cyanidin‐3‐O‐glucoside (C3G) is one of the principal types of anthocyanins. To understand the interaction between C3G and bovine serum albumin (BSA), fluorescence spectroscopy, ultraviolet–visible absorption, Fourier transform infrared spectroscopy, circular dichroism and molecular modeling techniques were used. Binding constant (Ka) and the number of binding sites (n) were calculated. The quenching mechanism of fluorescence of BSA by C3G was discussed. The results studied by Fourier transform infrared spectroscopy and circular dichroism experiments indicate that the secondary structures of the protein have been changed by the interaction of C3G with BSA. The result of molecular modeling confirmed that the C3G bound to the site I (sub‐domain IIA) of BSA, and that the hydroxyl groups in the B ring of C3G took part in the binding with BSA. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

10.
Diamine‐sarcophagine (DiAmsar) binding to human serum albumin (HSA) and bovine serum albumin (BSA) was investigated under simulative physiological conditions. Fluorescence spectra in combination with Fourier transform infrared (FT‐IR), UV‐visible (UV–vis) spectroscopy, cyclic voltammetry (CV), and molecular docking method were used in the present work. Experimental results revealed that DiAmsar had an ability to quench the HSA and BSA intrinsic fluorescence through a static quenching mechanism. The Stern–Volmer quenching rate constant (Ksv) was calculated as 0.372 × 103 M‐1 and 0.640 × 103 M‐1 for HSA and BSA, respectively. Moreover, binding constants (Ka), number of binding sites (n) at different temperatures, binding distance (r), and thermodynamic parameters (?H°, ?S°, and ?G°) between DiAmsar and HSA (or BSA) were calculated. DiAmsar exhibited good binding propensity to HSA and BSA with relatively high binding constant values. The positive ?H° and ?S° values indicated that the hydrophobic interaction is main force in the binding of the DiAmsar to HSA (or BSA). Furthermore, molecular docking results revealed the possible binding site and the microenvironment around the bond. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

11.
l ‐glutathione capped highly fluorescent CdTe quantum dots (QDs) were prepared by an aqueous approach and used as fluorescent labels to link albumin bovine serum (BSA) and rat anti‐mouse CD4, which was expressed on mouse T‐lymphocyte and mouse spleen tissue. The sharp and narrow emission peaks showed that the as‐prepared QDs have desirable dispersibility, uniformity and good fluorescence properties. Both CdTe–BSA and CdTe–CD4 conjugates showed an enhancement of fluorescence intensity over that of bare CdTe QDs. The experimental result of gel electrophoresis confirmed the successful conjugation of CdTe–BSA and CdTe–CD4. The fluorescent microscopic images of CdTe–CD4 labeled mouse T‐lymphocyte cells and mouse spleen tissue were compared with that obtained from fluorescein isothiocyanate labeling. It was demonstrated that the CdTe QDs‐based probe exhibited much better photostability and fluorescence intensity than fluorescein isothiocyanate, showing a good application potential in the immuno‐labeling of cells and tissues. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

12.
The influence of albumin towards the metabolism behavior of fenoprofen enantiomers and relevant drug–drug interaction was investigated in the present study. The metabolic behavior of fenoprofen enantiomers was compared in a phase II metabolic incubation system with and without bovine serum albumin (BSA). BSA supplement increased the binding affinity parameter (Km) of (R)‐fenoprofen towards human liver microsomes (HLMs) from 148.3 to 214.4 μM. In contrast, BSA supplement decreased the Km of (S)‐fenoprofen towards HLMs from 218.2 to 123.5 μM. For maximum reaction velocity (Vmax), the addition of BSA increased the Vmax of (R)‐fenoprofen from 1.3 to 1.6 nmol/min/mg protein. In the contrast, BSA supplement decreased the Vmax value from 3.3 to 1.5 nmol/min/mg protein. Andrographolide–fenoprofen interaction was used as an example to investigate the influence of BSA supplement towards fenoprofen‐relevant drug–drug interaction. The addition of 0.2% BSA in the incubation system significantly decreased the inhibition potential of andrographolide towards (R)‐fenoprofen metabolism (P < 0.001). Different from (R)‐fenoprofen, the addition of BSA significantly increased the inhibition potential of andrographolide towards the metabolism of (S)‐fenoprofen. BSA supplement also changed the inhibition kinetic type and parameter of andrographolide towards the metabolism of (S)‐fenoprofen. In conclusion, albumin supplement changes the metabolic behavior of fenoprofen enantiomers and the fenoprofen–andrographolide interaction. Chirality 27:436–440, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

13.
Spectrofluoremetric technique was employed to study the binding behavior of hydralazine with bovine serum albumin (BSA) at different temperatures. Binding study of bovine serum albumin with hydralazine has been studied by ultraviolet–visible spectroscopy, fluorescence spectroscopy and confirmed by three‐dimensional, synchronous, circular dichroism, and Raman spectroscopic methods. Effect of β‐cyclodextrin on binding was studied. The experimental results showed a static quenching mechanism in the interaction of hydralazine with bovine serum albumin. The binding constant and the number of binding sites are calculated according to Stern–Volmer equation. The thermodynamic parameters ?Ho, ?Go, ?So at different temperatures were calculated. These indicated that the hydrogen bonding and weak van der Waals forces played an important role in the interaction. Based on the Förster's theory of non‐radiation energy transfer, the binding average distance, r, between the donor (BSA) and acceptor (hydralazine) was evaluated and found to be 3.95 nm. Spectral results showed that the binding of hydralazine to BSA induced conformational changes in BSA. The effect of common ions on the binding of hydralazine to BSA was also examined. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

14.
Microsporidia are obligate intracellular parasites with extremely reduced genomes and a dependence on host‐derived ATP. The microsporidium Encephalitozoon cuniculi proliferates within a membranous vacuole and we investigated how the ATP supply is optimized at the vacuole–host interface. Using spatial EM quantification (stereology), we found a single layer of mitochondria coating substantial proportions of the parasitophorous vacuole. Mitochondrial binding occurred preferentially over the vegetative ‘meront’ stages of the parasite, which bulged into the cytoplasm, thereby increasing the membrane surface available for mitochondrial interaction. In a broken cell system mitochondrial binding was maintained and was typified by electron dense structures (< 10 nm long) bridging between outer mitochondrial and vacuole membranes. In broken cells mitochondrial binding was sensitive to a range of protease treatments. The function of directly bound mitochondria, as measured by the membrane potential sensitive dye JC‐1, was indistinguishable from other mitochondria in the cell although there was a generalized depression of the membrane potential in infected cells. Finally, quantitative immuno‐EM revealed that the ATP‐delivering mitochondrial porin, VDAC, was concentrated atthe mitochondria‐vacuole interaction site. Thus E. cuniculi appears to maximize ATP supply by direct binding of mitochondria to the parasitophorous vacuole bringing this organelle within 0.020 microns of the growing vegetative form of the parasite. ATP‐delivery is further enhanced by clustering of ATP transporting porins in those regions of the outer mitochondrial membrane lying closest to the parasite.  相似文献   

15.
Anthocyanin is one of the flavonoid phytopigments with specific health benefits. The interaction between delphinidin‐3‐O‐glucoside (D3G) and bovine serum albumin (BSA) was investigated by fluorescence spectroscopy, synchronous fluorescence spectroscopy, three‐dimensional fluorescence spectroscopy, ultraviolet‐visible absorption spectroscopy, circular dichroism spectroscopy and molecular modeling. D3G effectively quenched the intrinsic fluorescence of BSA via static quenching. The number of binding sites and binding constant Ka were determined, and the hydrogen bonds and van der Waals forces played major roles in stabilizing the D3G–BSA complex. The distance r between donor and acceptor was obtained as 2.81 nm according to Förster's theory. In addition, the effects of pH and metal ions on the binding constants were discussed. The results studied by synchronous fluorescence, three‐dimensional fluorescence and circular dichroism experiments indicated that the secondary structures of the protein has been changed by the addition of D3G and the α‐helix content of BSA decreased (from 56.1% to 52.4%). Furthermore, the study of site marker competitive experiments and molecular modeling indicated that D3G could bind to site I of BSA, which was in the large hydrophobic cavity of subdomain IIA. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

16.
Natural substances have been attracted several researchers in the recent years, because of its potential antioxidant, anti‐inflammatory and anti‐cancer properties. We have investigated the effect of carnosine on cell viability, apoptosis, DNA damage, reactive oxygen species (ROS) and caspase 3 enzyme expression in human cervical carcinoma and Madin‐Darby Kidney Cells (MDCK) cells . Carnosine inhibited cancer cell growth up to 23%. ROS level was increased up to 30 and 31% in MDCK and HeLa cells respectively. Tunnel assay showed 42 and 14% of positive apoptotic cells in cancer and normal cells respectively. The alteration in mitochondrial and nuclear morphology was determined. The extended lace‐like network of normal mitochondria found in control cells. Carnosine treatment significantly altered the mitochondrial morphology of normal cervical carcinoma cell. Mitochondria were condensed clump structures in carnosine treated cancer cells. Carnosine reduced the number of colonies of cervical carcinoma cells. Caspase 3 expression was corresponded to the appearance of immunofluorescence in the cytoplasm. Caspase 3 expression was gradually increased in cervical carcinoma cells. In Silico, docking study was performed to recognize the binding activity of carnosine against a subunit of the caspase 3 , and carnosine was able to bind to the drug binding pocket of caspase 3. The glide energy is ?5.2 kcal/mol, suggesting the high binding affinity of carnosine to caspase 3. Taking all these data together, the natural dipeptide L‐carnosine could be a suitable antiproliferative agent in cervical carcinoma cells. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

17.
Oleanane‐type triterpenoids serve as an important group of plant secondary metabolites with a variety of biological activities and the C‐3 position substitution pattern is a significant structural feature for their biological activities. Three selected oleanane‐type triterpenoids (glycyrrhizin, glycyrrhetinic acid, and carbenoxolone) bearing different substituents (glucuronic acid dimer, hydroxyl, and succinyl groups) at the C‐3 position were studied for their affinities to bind bovine serum albumin (BSA) by steady‐state fluorescence, synchronous, three‐dimensional fluorescence and ultraviolet–visible (UV–vis) absorption spectra. The binding mechanism of the triterpenoids to BSA is due to the formation of the triterpenoids–BSA complex and the binding affinity is strongest for carbenoxolone and ranked in the order carbenoxolone > glycyrrhetinic acid > glycyrrhizin. The thermodynamic parameters calculated at different temperatures showed that triterpenoids binding to BSA primarily depended on hydrophobic interaction and hydrogen bonding. The distance between the bound triterpenoid and BSA was determined on the basis of the Förster's energy transfer theory. Displacement experiments using phenylbutazone and ibuprofen showed the binding site of triterpenoids on BSA at subdomain IIA (Sudlow's site I). The effect of triterpenoids on BSA conformation was analyzed by UV–vis absorption, and synchronous and three‐dimensional fluorescence spectra. These results revealed that the C‐3 position substitution pattern significantly affects the structure–affinity relationships of oleanane‐type triterpenoid binding to BSA and further affects the bioavailability of triterpenoids in the blood circulatory system. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

18.
In the present investigation, the protein‐binding properties of naphthyl‐based hydroxamic acids (HAs), N‐1‐naphthyllaurohydroxamic acid ( 1 ) and N‐1‐naphthyl‐p‐methylbenzohydroxamic acid ( 2 ) were studied using bovine serum albumin (BSA) and UV–visible spectroscopy, fluorescence spectroscopy, diffuse reflectance spectroscopy–Fourier transform infrared (DRS–FTIR), circular dichroism (CD), and cyclic voltammetry along with computational approaches, i.e. molecular docking. Alteration in the antioxidant activities of compound 1 and compound 2 during interaction with BSA was also studied. From the fluorescence studies, thermodynamic parameters such as Gibb's free energy (ΔG), entropy change (ΔS) and enthalpy change (ΔH) were calculated at five different temperatures (viz., 298, 303, 308, 313 or 318 K) for the HAs–BSA interaction. The results suggested that the binding process was enthalpy driven with dominating hydrogen bonds and van der Waals’ interactions for both compounds. Warfarin (WF) and ibuprofen (IB) were used for competitive site‐specific marker binding interaction and revealed that compound 1 and compound 2 were located in subdomain IIA (Sudlow's site I) on the BSA molecule. Conclusions based on above‐applied techniques signify that various non‐covalent forces were involved during the HAs–BSA interaction. Therefore the resulted HAs–BSA interaction manifested its effect in transportation, distribution and metabolism for the drug in the blood circulation system, therefore establishing HAs as a drug‐like molecule.  相似文献   

19.
Bovine serum albumin (BSA) and anti‐BSA polyclonal antibody were used as model polypeptides to examine the movement of foreign proteins across the insect digestive system and their accumulation in hemolymph of fourth stadium tobacco budworms, Heliothis virescens (Fabricius) (Lepidoptera: Noctuidae). Hydrateable meal pads were developed in these studies as a method for easily introducing compounds into the insect digestive system. When insects were allowed to feed continuously on hydrated meal pads containing 0.8 mg of anti‐BSA per gram diet, the level of antibody found in hemolymph was 2.4 ± 0.1 and 3.4 ± 0.1 µg ml?1 (average  1 SEM) after 8 and 16 h, respectively, as determined by enzyme‐linked immunosorbant assay (ELISA). Continuous feeding on hydrated meal pads containing the same concentration of BSA produced hemolymph concentrations of 1.5 ± 0.1 and 1.6 ± 0.1 µg ml?1 hemolymph at 8 and 16 h, respectively. Western blot analyses demonstrated that BSA and anti‐BSA both retained their primary and multimeric structure and that anti‐BSA maintained its antigenic activity in the meal pads and after movement from meal pads into the hemolymph. When 1 µg of anti‐BSA or BSA was injected into the hemocoel of fourth instars, the concentrations decreased with time and 120 min after injection were 20% and 0.6% of the original concentration, respectively. When added at the same concentration to plasma in vitro, the decrease was 81.5% and 57.5%, respectively, at 2 h. The accumulation of native anti‐BSA and BSA protein in insect hemolymph is the result of their rate of movement across the gut and their rate of turnover in hemolymph. Movement of anti‐BSA and BSA across the digestive system was also noted in Helicoverpa zea (Boddie) (Lepidoptera: Noctuidae), Acheta domesticus (L.) (Orthoptera: Gryllidae), and Gromphadorhina portentosa (Schaum) (Blattaria: Blattellidae). Anti‐BSA and BSA were not detected in the hemolymph of Manduca sexta (L.) (Lepidoptera: Sphingidae) after feeding.  相似文献   

20.
We describe the formulation of bovine serum albumin nanoparticles (BSA‐NPs) by the coacervation method using surfactants. Plasmids (pUC18, pUC18egfp and pBBR1MCS‐2) isolated from E. coli were incorporated into the BSA matrix by incubating in albumin solution prior to formulation of NPs. Plasmid incorporation was calculated by % yield, entrapment efficiency, DNA loading capacity and release of entrapped DNA by comparing with blank NPs. BSA‐DNA binding studies were carried out by using fluorescence spectroscopy and Fourier Transform Infra Red Spectroscopy (FT‐IR). The surface charge distribution of the NPs loaded with plasmid was calculated using zeta potential. The photoluminescence of BSA‐NPs was quenched when loaded with pDNA, confirming the interaction of DNA with BSA. Altogether, these results provide evidences for the excellent DNA carrying efficiency of BSA‐NPs without loss of plasmid's integrity. The NPs were used to transfect E. coli DH5α strain lacking ampicillin resistance. They, however, showed ampicillin resistance subsequent to transfection with plasmid encoding ampicillin resistance gene. Effect of transfection was confirmed by confocal microscopy and by the isolation of the plasmid by agarose gel electrophoresis from the transfected bacterial culture. This study clearly demonstrates the efficacy of BSA‐NPs as delivery vehicle for pDNA transfection. Copyright © 2014 John Wiley & Sons, Ltd.  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