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 15 毫秒
1.
The noncovalent complexes of five cytotoxic protoberberine alkaloids, that is, berberine, palmatine, jatrorrhizine, coptisine, and berberrubine with several double-stranded oligodeoxynucleotides were systematically investigated by using electrospray ionization mass (ESI-MS) and fluorescence spectrometric methods, with the aim of establishing the structure-activity relationships. ESI-MS spectrometric studies indicated that these five alkaloids showed both 1:1 and 1:2 binding stoichiometries with d(AAGAATTCTT)(2), d(AAGGATCCTT)(2), and d(AAGCATGCTT)(2). Their relative binding affinities toward these three double-stranded DNA were semi-quantitatively evaluated by measuring the ratios of the complex signals ([ds+alkaloid-5H](4-)+[ds+2alkaloid-6H](4-)) to those of the duplexes ([ds-4H](4-)) and also by ESI-MS competitive binding experiments. These experiments established the relative binding affinities of five protoberberine alkaloids in the order of palmatine>jatrorrhizine>coptisine>berberine>berberrubine with d(AAGAATTCTT)(2), palmatinecoptisine>jatrorrhizineberberine>berberrubine with d(AAGGATCCTT)(2) and palmatine>jatrorrhizinecoptisine>berberine>berberrubine with d(AAGCATGCTT)(2). Significantly, these alkaloids except berberrubine bound to d(AAGGATCCTT)(2) and d(AAGCATGCTT)(2) with the affinities comparable to Hoechst 33258, a typical DNA minor groove binder. The relative binding preferences of berberine, palmatine, and coptisine with these three double-stranded DNA were further quantitatively assessed by their association constants obtained from fluorescence titration experiments. The values revealed the order of relative binding affinities as berberine>coptisine>palmatine with d(AAGAATTCTT)(2) and coptisine>berberine>palmatine with d(AAGGATCCTT)(2) and d(AAGCATGCTT)(2). These results were not in full agreement with those obtained from ESI-MS experiments, maybe due to the different measuring solution conditions. The results from ESI-MS and fluorescence titration experiments indicated that the sequence selectivities of these five alkaloids were not significant and remarkable AT- or GC-rich DNA binding preferences were not obtained, in contrast to the report that berberine binds preferentially to AT-rich DNA. To provide further insight into the sequence selectivities, the association constants of berberine with d(AAGATATCTT)(2), 5'-AAGTAATCTT-3'/5'-AAGATTACTT-3', d(AAGGGCCCTT)(2), d(AAGGCGCCTT)(2), and 5'-AAGGCCGCTT-3'/5'-AAGCGGCCTT-3', that is double helical DNA from AT-rich to GC-rich sequences, were further measured by fluorescence titration methods. No significant differences in their association constants were observed, suggesting that berberine showed no remarkable sequence selectivities.  相似文献   

2.
Six novel berberine dimers (3a-f) were synthesized in 37-84% yield from the reaction of berberrubine (2) with dihaloalkanes of varying lengths from two to seven carbons. Their interactions with calf thymus (CT) DNA and three double helical oligodeoxynucleotides, d(AAGAATTCTT)2, d(AAGCATGCTT)2, and d(TAAGAATTCTTA)2, were investigated by means of fluorometric titration and ethidium bromide (EB) displacement experiments. Compared with the monomeric parent berberine (1), these dimers' DNA-binding affinities increased up to approximately 100-fold, suggesting a cooperative interaction of the two berberine subunits in the molecules. Furthermore, these dimers linked by different spacers show a prominent structure-activity relationship when bound with oligodeoxynucleotides. The relative binding affinities are in the order of 3b>3a>3c>3d>3e>3f with d(AAGAATTCTT)2 and d(TAAGAATTCTTA)2, and 3b>3c>3a>3d>3e>3f with d(AAGCATGCTT)2. Dimer 3b, linked with a propyl chain, exhibits the highest binding affinity. This suggests that a propyl chain may be the most suitable spacer to bridge the two berberine units for DNA binding. Spectrophotometric titration and competitive EB displacement of berberine (1) and dimer 3b indicate that both berberine and its dimers form intercalating complexes with duplex DNA. A larger redshift, a stronger hypochromic effect, and a much higher EB displacement ratio, observed in 3b, indicate that the dimer is in more intimate contact with DNA than berberine. In addition, no obvious binding of canadine (4), a hydrogenated product of berberine, with CT DNA was observed, suggesting critical roles of the quaternary ammonium cation and planar structure in the DNA-binding of berberine.  相似文献   

3.
Six jatrorrhizine homodimers and berberine-jatrorrhizine heterodimers have been synthesized in moderate to good yields from the reaction of jatrorrhizine with alpha,omega-dibromoalkanes and 9-O-(omega-bromoalkyl)berberines, respectively. Their binding activities toward calf thymus (CT) DNA and three double-stranded oligodeoxynucleotides, d(AAGAATTCTT)(2), d(TAAGAATTCTTA)(2), and d(TTAAGAATTCTTAA)(2), were investigated by means of spectrofluorimetric and spectrophotometric titrations. The results indicate that these dimers exhibit enhanced DNA-binding affinities due to the cooperative interaction of the two protoberberine subunits. A comparative study of the DNA-binding behaviors of berberine homodimers, jatrorrhizine homodimers, and berberine-jatrorrhizine heterodimers suggests that spacer length and attaching position are of great importance in modulating their DNA-binding affinities.  相似文献   

4.
This communication describes the facile synthesis of five novel berberine dimers and their strong affinities toward double-stranded DNA. These berberine dimers were synthesized in 37-84% yields from the reaction of berberrubine with dihaloalkanes of varying lengths, and fully characterized by HRMS and 1H NMR. Compared with the monomeric parent berberine, these dimers showed greatly enhanced binding affinities up to approximately 100-fold, with two double helical oligodeoxynucleotides, d(AAGAATTCTT)2 and d(TAAGAATTCTTA)2, which was investigated by means of fluorescence spectrometry.  相似文献   

5.
Berberine, an isoquinoline plant alkaloid, belongs to the structural class of protoberberines. Recently, the ability of these compounds to act as Topoisomerase I or II poisons, was related to the antitumor activity. The binding of protoberberins to DNA has been studied and the partial intercalation into the double helix has been considered responsible for their activity. We have studied the interaction of berberine with the double helix oligonucleotides d(AAGAATTCTT)(2), d(GCGATCGC)(2), d(CGTATACG)(2), d(CGTACG)(2), 5'-d(ACCTTTTTGATGT)-3'/5(ACATCAAAAAGGT)-3' and with the single strand 5'-d(ACATCAAAAAGGT)-3', by 1H, 31P NMR and UV spectroscopy. Phosphorus resonance experiments were performed to detect small conformational changes of the phosphoribose backbone, in the case that an intercalation process occurs. Our data reveal that berberine does not intercalate into the duplexes studied, and binds preferentially to AT rich sequences. The structure of the complex with d(AAGAATTCTT)(2) was determined by using proton 2D NOESY spectra, which allowed to obtain several NOE contacts between the drug and the nucleotide. Structural models were built up by Molecular Mechanics (MM) and Molecular Dynamics (MD) calculations, by using the inter-proton distances derived from the NOE values. Berberine results to be located in the minor groove, lying with the convex side on the helix groove and presenting the positively charged nitrogen atom close to the negative ionic surface of the oligomer. The large 1H chemical shifts variation, observed for the drug when it is added to the above duplexes, as well as to the single strand oligomer, was interpreted with non-specific ionic interactions. The binding constants were measured by UV and NMR spectroscopy. They are strongly affected by the ionic strength and by the self-association process, which commonly occurs with this type of drugs. A dimerisation constant was measured and the value was included in the calculations of the binding constants. The results obtained show that the non-specific ionic interactions represent the major contribution to the values of the binding constants. These parameters, as well as the protons chemical shift variation of the ligand, are thus not diagnostic for the identification of a drug/DNA complex.  相似文献   

6.
Lah J  Vesnaver G 《Biochemistry》2000,39(31):9317-9326
Circular dichroism (CD), isothermal calorimetric titrations (ITC), and temperature-dependent UV spectroscopy were used to investigate binding of the minor groove-directed ligands distamycin A (Dst) and netropsin (Net) to the following duplexes: d(GTTAGTATTTGG). d(CCAAATACTAAC), d(GTTAGTATATGG).d(CCATATACTAAC), d(GTTAGTACTTGG). d(CCAAGTACTAAC), and d(GTTAGTAGTTGG).d(CCAACTACTAAC). Our results reveal that Dst binds within the minor grooves of these dodecamers that contain five-AT and/or four-AT.GC binding sites exclusively in a dimeric high-affinity 2:1 binding mode (K approximately 10(16) M(-)(2)). By contrast, Net exhibits high-affinity binding only when it binds in a 1:1 mode (K(1) approximately 10(9) M(-)(1)) to the two duplexes that contain five-AT sites (5'-TATTT-3' and 5'-TATAT-3'). Its further binding to these two duplexes occurs in a low-affinity mode (K(2) approximately 10(6) M(-)(1)) and results in the formation of 2:1 Net-DNA complexes. To the other two duplexes that contain sequences with at most three AT consecutive base pairs Net binds in two distinctive low-affinity 1:1 binding modes (K(1) approximately 10(7) M(-)(1), K(2) approximately 10(6) M(-)(1)). Competition experiments (CD and ITC titrations) reveal that Dst entirely displaces Net from its 1:1 and 2:1 complexes with any of the four duplexes. We discuss and interpret our optical and calorimetric results in the context of the available structural information about the complexes between DNA and the sequence-specific minor groove binders Dst and Net.  相似文献   

7.
This study examines the characteristics of binding of berberine to the human telomeric d[AG(3)(T(2)AG(3))(3)] quadruplex. By employing UV-visible spectroscopy, fluorescence spectroscopy and isothermal titration calorimetry, we found that the binding affinity of berberine to the human telomeric quadruplex is 10(6). The complete thermodynamic profile for berberine binding to the quadruplex, at 25 degrees C, shows a small negative enthalpy (DeltaH) of -1.7 kcal.mol(-1), an entropy change with TDeltaS of +6.5 kcal.mol(-1), and an overall favorable free energy (DeltaG) of -8.2 kcal.mol(-1) .Through the temperature dependence of DeltaH, we obtained a heat capacity (DeltaC(p)) of -94 (+/- 5) cal.mol(-1).K(-1). The osmotic stress method revealed that there is an uptake of 13 water molecules in the complex relative to the free reactants. Furthermore, the molecular modeling studies on different quadruplex-berberine complexes show that berberine stacking at the external G-quartet is mainly aided by the pi-pi interaction and the stabilization of the high negative charge density of O6 of guanines by the positively charged N7 of berberine. The theoretical heat capacity (DeltaC(p)) values for quadruplex-berberine models are -89 and -156 cal.mol(-1).K(-1).  相似文献   

8.
The interaction of berberine and its 9-substituted derivatives with human telomeric DNA d[G(3)(T(2)AG(3))(3)](telo21) has been investigated via CD spectroscopy, fluorescence spectroscopy, PCR-stop assay, competitive dialysis, and telomerase repeat amplification protocol (TRAP) assay. The results indicated that these semisynthesized compounds could induce and stabilize the formation of anti-parallel G-quadruplex of telomeric DNA in the presence or absence of metal cations. Compared with berberine, the 9-substituted derivatives exhibit stronger binding affinity with G-quadruplex and higher inhibitory activity for telomerase. Introduction of a side chain with proper length of methylene and terminal amino group to the 9-position of berberine would significantly strengthen the binding affinity with G-quadruplex, resulting in increasing inhibitory effects on the amplification of telo21 DNA and on the telomerase activity.  相似文献   

9.
A variety of organic cations, cationic lipids, low molecular weight alcohols, sodium dodecylsulfate, trehalose, glycerol, low molecular weight polyethylene glycols, and DMSO were tested for their ability to modulate the stability of the triplexes d(C(+)-T)(6):d(A-G)(6);d(C-T)(6), d(T)(21):d(A)(21);d(T)(21), poly r(U:A U) and their respective core duplexes, d(A-G)(6);d(C-T)(6), d(A)(21);d(T)(21), poly r(A-U). Very substantial enhancement of triplex stability over that in a physiological salt buffer at pH 7 is obtained with different combinations of triplex and high concentrations of these additives, e.g. trimethylammonium chloride and d(C(+)-T)(6):d(A-G)(6);d(C-T)(6); 2-propanol and d(T)(21):d(A)(21);d(T)(21); ethanol and poly r(U:A;U). Triplex formation is even observed with a 1:1 strand mixture of d(A-G)(6) and d(C-T)(6) in the presence of dimethylammonium, tetramethylammonium, and tetraethylammonium-chloride, as well as methanol, ethanol, and 2-propanol. Triplex stability follows the water structure-making ability (and in some cases the duplex unwinding ability) of the organic cations, the low molecular weight alcohols and other neutral organic compounds, whereas water structure-breaking additives decrease triplex stability. These findings are consistent with those reported in the accompanying paper that triplex formation occurs with a net uptake of water. Since the findings suggest that third strand-binding is facilitated by unwinding of the target duplex, it is inferred that triplex formation may be enhanced by nucleic acid binding proteins operating similarly.  相似文献   

10.
目的:观察小檗碱对流感病毒感染所致病毒性肺炎小鼠肺血管通透性的影响,并探讨其作用机制。方法:BALB/c小鼠108只随机分为3组,正常组、模型组、小檗碱组,25μL 50LD50病毒液滴鼻建立流感病毒感染的小鼠肺炎模型,感染后1 h,正常组和模型组予以双蒸水灌胃,小檗碱组予药物0.005 g.kg-1d-1腹腔注射;各组均给药2次/d,连续给药5 d。感染后的2 d、4 d、6 d,处死小鼠,肺组织称重以检测肺含水量;1%伊文氏兰5 mL/kg尾静脉注射检测肺血管通透性;Bicinchoninic acid(BCA)法检测肺泡灌洗液(BALF)中蛋白含量;放免法或酶免法测定肺组织中PGE2、PLA2及LT-B4含量。结果:病毒感染后,模型组肺含水量持续升高,肺血管通透性及BALF蛋白含量在感染后第4天开始明显升高,小檗碱降低了肺含水量、肺血管通透性及BALF蛋白含量(P<0.01);模型组肺组织中PGE2、PLA2、LT-B4的含量明显升高,小檗碱不同程度地抑制了PGE2、PLA2、LT-B4的表达。结论:小檗碱通过抑制流感病毒感染后肺组织中PGE2、PLA2、LT-B4的释放,降低了肺血管通透性及肺含水量,对病毒性肺炎中肺水肿的形成,起到一定的治疗作用。  相似文献   

11.
Valente EJ  Moore MC 《Chirality》2000,12(1):16-25
Resolution of the isomeric 2'-, 3'-, and 4'-fluoromandelic acids with (+)-(1S;2S)-pseudoephedrine in 95% ethanol produces both well and poorly discriminating, hydrated and unsolvated binary salts. Seven observed diastereomeric phases are represented by five crystal structure types including three of the four types observed in the pseudoephedrine mandelates. Type a: monoclinic hemihydrate less-soluble (L) (R)-3'-fluoromandelate and more-soluble (M) (R)-4'-fluoromandelate (I); type b: orthorhombic unsolvated M (S)-2'-fluoromandelate; type c: orthorhombic unsolvated L (R)-2'-fluoromandelate; type d: orthorhombic dihydrate M (S)-3'-fluoromandelate and L (S)-4'-fluoromandelate; type e: monoclinic unsolvated M (R)-4'-fluoromandelate (II). Largest (15-fold) discriminating solubilities in 95% ethanol are found between the diastereomers with 2'-fluoromandelic acid, 50% more than in the corresponding ephedrine system. Principle interionic interactions are hydrogen-bonds between protonated secondary ammonium ions and carboxylates. Infinite chains of these are found in type c, with a four-atom repeating unit H-N(+)-H.O(-C(-)-O) [C(2)(1)(4)], and in types b and d, with a six-atom repeating unit H-N(+)-H.O-C(-)-O [C(2)(2)(6)]. Water of crystallization intervenes in the chains of type a but not of type d hydrated salts, according with higher average dehydration temperatures in the former. Hydrated salts in general are excessively soluble in 95% ethanol.  相似文献   

12.
Cancer stem cells play an important role in metastasis and the relapse of drug resistant cancers. Side-population (SP) cells are capable of effluxing Hoechst 33342 dye and are referred to as cancer stem cells. We investigated the effect of berberine on pancreatic cancer stem cells of PANC-1 and MIA PaCa-2. For both cell lines, the proportions of SP cells in the presence of berberine were investigated and compared to the proportions in the presence of gemcitabine, a standard pancreatic anti-cancer drug. The proportions of SP cells in the PANC-1 and MIA PaCa-2 cell lines were about 9 and <0.1 %, respectively. After berberine and gemcitabine treatments, the SP cell proportion of PANC-1 decreased to 5.7 ± 2.0 and 6.8 ± 0.8 %, respectively, which compares to the control proportion of (9.7 ± 1.7). After berberine and gemcitabine treatment of PANC-1, of the four stem cell-associated genes (SOX2, POU5F1, NANOG, and NOTCH1), all but NOTCH1 were down-regulated. Unfortunately, the effect of berberine and gemcitabine treatments on MIA PaCa-2 SP cells could not be clearly observed because SP cells represented only a very small proportion of MIA PaCa-2 cells. However, SOX2, POU5F1, and NANOG genes were shown to be effectively down-regulated in the MIA PaCa-2 cell line as a whole. Taken together, these results indicate that berberine is as effective at targeting pancreatic cancer cell lines as gemcitabine. Therefore, we believe that POU5F1, SOX2, and NANOG can serve as potential markers, and berberine may be an effective anti-cancer agent when targeting human pancreatic cancer cells and/or their cancer stem cells.  相似文献   

13.
The fluorescence enhancement of berberine hydrochloride (BBH) as a result of complex with β-cyclodextrin (β-CD) is investigated. The mechanism of the inclusion was studied and discussed by spectrofluoremetry and infrared spectrograms. The results showed that a 1∶1 (β-CD: BBH) complex was formed with an apparent association constant of 4.23×102 L/mol. Based on the enhancement of the fluorescent intensity of berberine hydrochloride, a new spectrofluorimetric method for the determination of BBH in the presence of β-CD was developed. The linear range was 1.00∼4.00 µg/mL with the detection limit of 5.54 ng/mL. The proposed method was successfully applied to the determination of BBH in tablets.  相似文献   

14.
Bifurcation is a basic form of vascular connection. It is composed of a parent vessel of diameter d(0), and two daughter vessels, d(1) and d(2), where d(0) > d(1) >/= d(2). Optimal values for the bifurcation area ratio, beta = (d(1)(2) + d(2)(2))/d(0)(2), and the junction exponent, x, in d(0)(x) = d(1)(x) + d(2)(x), are postulated to be universal in nature. However, we have hypothesized that the perinatal pulmonary arterial circulation is an exception. Arterial diameters were measured in pulmonary vascular casts of a fetal lamb (140 days gestation/145 days term) and a neonatal lamb (1 day old). The values for beta and x were evaluated in 10,970 fetal and 846 neonatal bifurcations sampled from the proximal and intermediate arterial regions. Mean values and confidence intervals (CI) for the fetus were beta = 0.890 (0.886-0.895 CI) and x = 1.75 (1.74-1.76 CI); and for the newborn were beta = 0.913 (0.90-0.93 CI) and x = 1. 79 (1.75-1.82 CI). These values are significantly different from Murray's law (beta > 1, x = 3) or the West-Brown-Enquist law (beta = 1, x = 2). Therefore, perinatal pulmonary bifurcation design appears to be distinctive and exceptional. The decreasing cross-sectional area with branching leads to the hemodynamic consequence of shear stress amplification. This structural organization may be important for facilitating vascular development at low flow rates; however, it may be the origin of unstable reactivity if elevated blood flow and pressure occurs.  相似文献   

15.
Berberine, an alkaloid originally extracted from the plant Coptis chinensis and other herb plants, has been used as a pharmacological substance for many years. The therapeutic effect of berberine has been attributed to its interaction with nucleic acids and blocking cell division. However, levels of berberine entering individual microbial cells minimal for growth inhibition and its effects on bacterial spores have not been determined. In this work the kinetics and levels of berberine accumulation by individual dormant and germinated spores were measured by laser tweezers Raman spectroscopy and differential interference and fluorescence microscopy, and effects of berberine on spore germination and outgrowth and spore and growing cell viability were determined. The major conclusions from this work are that: (1) colony formation from B. subtilis spores was blocked ~ 99% by 25 μg/mL berberine plus 20 μg/mL INF55 (a multidrug resistance pump inhibitor); (2) 200 μg/mL berberine had no effect on B. subtilis spore germination with L-valine, but spore outgrowth was completely blocked; (3) berberine levels accumulated in single spores germinating with ≥ 25 μg/mL berberine were > 10 mg/mL; (4) fluorescence microscopy showed that germinated spores accumulated high-levels of berberine primarily in the spore core, while dormant spores accumulated very low berberine levels primarily in spore coats; and (5) during germination, uptake of berberine began at the time of commitment (T1) and reached a maximum after the completion of CaDPA release (Trelease) and spore cortex lysis (Tlysis).  相似文献   

16.
Payne JC  Rous BW  Tenderholt AL  Godwin HA 《Biochemistry》2003,42(48):14214-14224
Zinc binding to the two Cys(4) sites present in the DNA-binding domain (DBD) of nuclear hormone receptor proteins is required for proper folding of the domain and for protein activity. By utilizing Co(2+) as a spectroscopic probe, we have characterized the metal-binding properties of the two Cys(4) structural zinc-binding sites found in the DBD of human estrogen receptor alpha (hERalpha-DBD) and rat glucocorticoid receptor (GR-DBD). The binding affinity of Co(2+) to the two proteins was determined relative to the binding affinity of Co(2+) to the zinc finger consensus peptide, CP-1. Using the known dissociation constant of Co(2+) from CP-1, the dissociation constants of cobalt from hERalpha-DBD were calculated: K(d1)(Co) = 2.2 (+/- 1.0) x 10(-7) M and K(d2)(Co) = 6.1 (+/- 1.5) x 10(-7) M. Similarly, the dissociation constants of Co(2+) from GR-DBD were calculated: K(d1)(Co) = 4.1 (+/- 0.6) x 10(-7) M and K(d2)(Co) = 1.7 (+/- 0.3) x 10(-7) M. Metal-binding studies conducted in which Zn(2+) displaces Co(2+) from the metal-binding sites of hERalpha-DBD and GR-DBD indicate that Zn(2+) binds to each of the Cys(4) metal-binding sites approximately 3 orders of magnitude more tightly than Co(2+) does: the stoichiometric dissociation constants are K(d1)(Zn) = 1 (+/- 1) x 10(-10) M and K(d2)(Zn) = 5 (+/- 1) x 10(-10) M for hERalpha-DBD and K(d1)(Zn) = 2 (+/- 1) x 10(-10) M and K(d2)(Zn) = 3 (+/- 1) x 10(-10) M for GR-DBD. These affinities are comparable to those observed for most other naturally occurring structural zinc-binding sites. In contrast to the recent prediction by Low et. al. that zinc binding in these systems should be cooperative [Low, L. Y., Hernández, H., Robinson, C. V., O'Brien, R., Grossmann, J. G., Ladbury, J. E., and Luisi, B. (2002) J. Mol. Biol. 319, 87-106], these data suggest that the zincs that bind to the two sites in the DBDs of hERalpha-DBD and GR-DBD do not interact.  相似文献   

17.
The synthesis and characterization of the diiron(II) complex [Fe(2)(μ-OTf)(2)-(PIC(2)DET)(2)](BARF)(2) (2), where PIC(2)DET is a 2,3-diethynyltriptycene-linked dipicolinic methyl ester ligand, are described. The dication in 2, contains, [Fe(2)(μ-OTf)(2)(PIC(2)DET)(2)](2+) two symmetry-equivalent iron atoms with octahedral coordination geometries. Each metal ion has a N(2)O(4) atom donor set that includes four atoms from two picolinic ester N,O chelate rings, as well as two oxygen atoms from the bridging trifluoromethanesulfonate groups. The Fe(2)(μ-OTf)(2) core of 2 is stabilized by two PIC(2)DET ligands that bind the two metal ions in a head-to-head fashion, leading to an Fe···Fe distance of 5.173(1)?. Molar conductivity data for 2 are consistent with Fe(2)(μ-OTf)(2)(PIC(2)DET)(2)](2+) retaining its identity in acetone solutions, where it behaves as a 2:1 electrolyte. (1)H NMR spectroscopic, solution (d(6)-acetone) and solid-state magnetic susceptibility data all indicate that the iron atoms of 2 are high-spin (S = 2). A fit of the magnetic data (2 - 300K) to a spin-only isotropic exchange Hamiltonian H = -2JS(1)·S(2) are consistent with weak antiferromagnetic coupling between the two iron atoms with J ~ -0.99(2) cm(-1) and g = 2.10(1).  相似文献   

18.
19.
Berberine, a type of isoquinoline alkaloid isolated from Chinese medicinal herbs, has been reported to have various pharmacological activities. Studies have demonstrated that berberine has beneficial effects on vascular remodeling and alleviates restenosis after vascular injury. However, its mechanism of action on vascular smooth muscle cell migration is not fully understood. We therefore investigated the effect of berberine on human aortic smooth muscle cell (HASMC) migration. Boyden chamber assay was performed to show that berberine inhibited HASMC migration dosedependently. Real-time PCR and Western blotting analyses showed that levels of matrix metalloproteinase (MMP)-2, MMP-9, and urokinase-type plasminogen activator (u-PA) were reduced by berberine at both the mRNA and protein levels. Western blotting assay further confirmed that activities of c-Fos, c-Jun, and NF-κB were significantly attenuated. These results suggest that berberine effectively inhibited HASMC migration, possibly by down-regulating MMP-2, MMP-9, and u-PA; and interrupting AP-1 and NF-κB mediated signaling pathways. [BMB Reports 2014; 47(7): 388-392]  相似文献   

20.
Berberine is a plant alkaloid with anti-diabetic action. Activation of AMP-activated protein kinase (AMPK) pathway has been proposed as mechanism for berberine’s action. This study aimed to examine whether AMPK activation was necessary for berberine’s glucose-lowering effect. We found that in HepG2 hepatocytes and C2C12 myotubes, berberine significantly increased glucose consumption and lactate release in a dose-dependent manner. AMPK and acetyl coenzyme A synthetase (ACC) phosphorylation were stimulated by 20 µmol/L berberine. Nevertheless, berberine was still effective on stimulating glucose utilization and lactate production, when the AMPK activation was blocked by (1) inhibition of AMPK activity by Compound C, (2) suppression of AMPKα expression by siRNA, and (3) blockade of AMPK pathway by adenoviruses containing dominant-negative forms of AMPKα1/α2. To test the effect of berberine on oxygen consumption, extracellular flux analysis was performed in Seahorse XF24 analyzer. The activity of respiratory chain complex I was almost fully blocked in C2C12 myotubes by berberine. Metformin, as a positive control, showed similar effects as berberine. These results suggest that berberine and metformin promote glucose metabolism by stimulating glycolysis, which probably results from inhibition of mitochondrial respiratory chain complex I, independent of AMPK activation.  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