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.
We aimed to synthesize hydroxyethyl starch (HES) 200/0.5-loaded bovine serum albumin nanoparticles (HBNs) and investigate the compatibility and binding mechanism in simulated physiological environments. Here, to elucidate the morphology, biocompatibility, and formation mechanism of HBNs, techniques such as scanning electron microscopy, haemolysis test, fluorescence, and circular dichroism spectroscopy were applied. The thermodynamic parameters at body temperature (ΔS° = −26.7 J·mol−1·K−1, ΔH° = −3.20 × 104 J·mol−1, and ΔG = −2.35 × 104 J·mol−1) showed a 1:1 binding stoichiometry, which was formed by hydrogen bonds and van der Waals interactions. In addition, the conformational analysis showed that the microenvironment of fluorophores was altered with the adaptational protein secondary structural changes. Energy transfer occurred from the fluorophores to HES with a high possibility. All these results provided accurate and complete primary data for demonstrating the interaction mechanisms of HES with BSA, which helps to understand its pharmaceutical effects in blood.  相似文献   

2.
《Phytomedicine》2015,22(6):621-630
Background6-Shogaol, one of the main bioactive constituents of Zingiber officinale has been shown to possess various therapeutic properties. Interaction of a therapeutic compound with plasma proteins greatly affects its pharmacokinetic and pharmacodynamic properties.PurposeThe present investigation was undertaken to characterize the interaction between 6-shogaol and the main in vivo transporter, human serum albumin (HSA).MethodsVarious binding characteristics of 6-shogaol–HSA interaction were studied using fluorescence spectroscopy. Thermal stability of 6-shogaol–HSA system was determined by circular dichroism (CD) and differential scanning calorimetric (DSC) techniques. Identification of the 6-shogaol binding site on HSA was made by competitive drug displacement and molecular docking experiments.ResultsFluorescence quench titration results revealed the association constant, Ka of 6-shogaol–HSA interaction as 6.29 ± 0.33 × 104 M−1 at 25 ºC. Values of the enthalpy change (−11.76 kJ mol−1) and the entropy change (52.52 J mol−1 K−1), obtained for the binding reaction suggested involvement of hydrophobic and van der Waals forces along with hydrogen bonds in the complex formation. Higher thermal stability of HSA was noticed in the presence of 6-shogaol, as revealed by DSC and thermal denaturation profiles. Competitive ligand displacement experiments along with molecular docking results suggested the binding preference of 6-shogaol for Sudlow's site I of HSA.ConclusionAll these results suggest that 6-shogaol binds to Sudlow's site I of HSA through moderate binding affinity and involves hydrophobic and van der Waals forces along with hydrogen bonds.  相似文献   

3.
Interaction of procarbazine (PCZ) with calf thymus DNA was studied using biophysical and molecular docking studies. Procarbazine was to interact with DNA with a binding constant of 6.52 × 103 M−1 as calculated using ultraviolet‐visible spectroscopy. To find out the binding mode, molecular docking was performed that predicted PCZ to interact with DNA through groove binding mode with binding affinity of −6.7 kcal/mole. To confirm the groove binding nature, different experiments were performed. Dye displacement assays confirmed the non‐intercalative binding mode. Procarbazine displaced Hoechst dye from the minor groove of DNA while it was unable to displace intercalating dyes. There was no increase in the viscosity of DNA solution in presence of PCZ. Also, negligible change in the secondary structure of DNA was observed in presence of PCZ as evident by circular dichroism spectra. Procarbazine caused decrease in the melting temperature of DNA possibly because of decrease in the stability of DNA caused by groove binding interaction of PCZ with DNA.  相似文献   

4.
The present study was undertaken to gain insight into the associations of mercury(II) with dicysteinyl tripeptides in buffered media at pH 7.4. We investigated the effects of increasing the distance between cysteinyl residues on mercury(II) associations and complex formations. The peptide–mercury(II) formation constants and their associated thermodynamic parameters in 3-(N-morpholino)propanesulfonic acid (MOPS) buffered solutions were evaluated by isothermal titration calorimetry. Complexes formed in different relative ratios of mercury(II) to cysteinyl peptides in ammonium formate buffered solutions were characterized by LTQ Orbitrap mass spectrometry. The results from these studies show that n-alkyl dicysteinyl peptides (CP 14), and an aryl dicysteinyl peptide (CP 5) can serve as effective “double anchors” to accommodate the coordination sites of mercury(II) to form predominantly one-to-one Hg(peptide) complexes. The aryl dicysteinyl peptide (CP 5) also forms the two-to-two Hg2(peptide)2 complex. In the presence of excess peptide, Hg(peptide)2 complexes are also detected. Notably, increasing the distance between the ligating groups or “anchor points” in CP 15 does not significantly affect their affinity for mercury(II). However, the enthalpy change (ΔH) values (ΔH1  −91 kJ mol−1 and ΔH2  −66 kJ mol−1) for complex formation between CP 4 and 5 with mercury(II) are about one and a half times larger than the related values for CP 1, 2 and 3H1  −66 kJ mol−1 and ΔH2  46 kJ mol−1). The corresponding entropy change (ΔS) values (ΔS1  −129 J K−1 mol−1 and ΔS2  −116 J K−1 mol−1) of the structurally larger dicysteinyl peptides CP 4 and 5 are less entropically favorable than for CP 1, 2 and 3S1  −48 J K−1 mol−1 and ΔS2  −44 J K−1 mol−1). Generally, these associations result in a decrease in entropy, indicating that these peptide–mercury complexes potentially form highly ordered structures. The results from this study show that dicysteinyl tripeptides are effective in binding mercury(II) and they are promising motifs for the design of multi-cysteinyl peptides for binding more than one mercury(II) ion per peptide.  相似文献   

5.
The interaction of a recently certified kinase inhibitor Tofacitinib (TFB) with bovine serum albumin (BSA) has been studied, by spectroscopic and molecular docking studies. Spectrofluorimetric measurements at 3 different temperatures (288, 298, and 310 K) showed that TFB quench the intrinsic fluorescence of BSA upon forming a nonfluorescent complex. The intrinsic fluorescence data showed that TFB binds to BSA with binding constant (K b) of approximately 104M−1, affirming a significant affinity of TFB with BSA. The decrease in Stern‐Volmer quenching constant with increasing temperature exhibited the static mechanism of quenching. Negative value of ΔG (−6.94 ± 0.32 kcal·mol−1), ΔH (−7.87 ± 0.52 kcal·mol−1), and ΔS (−3.14 ± 0.42 cal·mol−1·K−1) at all 3 temperatures declared the reaction between BSA and TFB to be spontaneous and exothermic. Far‐UV circular dichroism spectroscopy results demonstrated an increase in helical content of BSA in the presence of TFB. Moreover, dynamic light scattering measurements showed that TFB resulted into a decrease in the hydrodynamic radii (from 3.6 ± 0.053 to 2.9 ± 0.02 nm) of BSA. Molecular docking studies confirmed that TFB binds near site II on BSA, hydrogen bonding, and hydrophobic interaction were involved in the BSA‐TFB complex formation. The present study characterizing the BSA‐TFB interaction could be significant towards gaining an insight into the drug pharmacokinetics and pharmacodynamics and also in the direction of rational drug designing with better competence, against emerging immune‐mediated diseases, ie, alopecia and rheumatoid arthritis.  相似文献   

6.
Many macromolecular interactions, including protein‐nucleic acid interactions, are accompanied by a substantial negative heat capacity change, the molecular origins of which have generated substantial interest. We have shown previously that temperature‐dependent unstacking of the bases within oligo(dA) upon binding to the Escherichia coli SSB tetramer dominates the binding enthalpy, ΔHobs, and accounts for as much as a half of the observed heat capacity change, ΔCp. However, there is still a substantial ΔCp associated with SSB binding to ssDNA, such as oligo(dT), that does not undergo substantial base stacking. In an attempt to determine the origins of this heat capacity change, we have examined by isothermal titration calorimetry (ITC) the equilibrium binding of dT(pT)34 to SSB over a broad pH range (pH 5.0–10.0) at 0.02 M, 0.2 M NaCl and 1 M NaCl (25°C), and as a function of temperature at pH 8.1. A net protonation of the SSB protein occurs upon dT(pT)34 binding over this entire pH range, with contributions from at least three sets of protonation sites (pKa1 = 5.9–6.6, pKa2 = 8.2–8.4, and pKa3 = 10.2–10.3) and these protonation equilibria contribute substantially to the observed ΔH and ΔCp for the SSB‐dT(pT)34 interaction. The contribution of this coupled protonation (∼ −260 to −320 cal mol−1 K−1) accounts for as much as half of the total ΔCp. The values of the “intrinsic” ΔCp,0 range from −210 ± 33 cal mol−1 °K−1 to −237 ± 36 cal mol−1K−1, independent of [NaCl]. These results indicate that the coupling of a temperature‐dependent protonation equilibria to a macromolecular interaction can result in a large negative ΔCp, and this finding needs to be considered in interpretations of the molecular origins of heat capacity changes associated with ligand‐macromolecular interactions, as well as protein folding. Proteins 2000;Suppl 4:8–22. © 2000 Wiley‐Liss, Inc.  相似文献   

7.
The binding of tiamulin with calf thymus DNA was systematically investigated using multispectroscopy and molecular modelling techniques. For DNA, once tiamulin was added, viscosity (η) and melting temperature (Tm) both exhibited an uptrend. The fluorescence performance of the tiamulin–DNA complex did not change with the ionic strength changes. The binding constant (Ka) of tiamulin for single-stranded DNA (ssDNA, 1.48 × 104 M−1) was obviously higher than that for double-stranded DNA (dsDNA, 9.51 × 103 M−1) at 291 K. The helix structure became looser and the base stack force became stronger for DNA due to the presence of tiamulin as seen from circular dichroic (CD) spectra. The intercalation binding mode of tiamulin with DNA was disclosed. Molecular modelling also revealed tiamulin inserting into the base pairs with the lowest binding free energy of −18.73 kJ mol−1 using van der Waals forces as well as hydrogen bonds.  相似文献   

8.
Interactions between efavirenz (EFZ) with calf thymus DNA (CT‐DNA) were investigated in vitro under stimulated physiological conditions using multispectroscopic techniques, cyclic voltammetry viscosity measurement, and gel electrophoresis. Methylene blue and acridine orange dyes were used as spectral probes by fluorescence spectroscopy. Hypochromicity was observed in ultra‐violet (UV) absorption band of EFZ. Considerable fluorescence enhancement of EFZ was observed in the presence of increasing amounts of DNA solution and the binding constants (Kf) and corresponding numbers of binding sites (n) were calculated at different temperatures. Thermodynamic parameters including enthalpy change (ΔH) and entropy change (ΔS) were calculated to be –304.78 kJ mol–1 and –924.52 J mol–1 K–1 according to the van ’t Hoff equation, which indicated that reaction is predominantly enthalpically driven. In addition, UV/vis absorption titration of DNA bases confirmed that EFZ interacted with guanine and cytosine preferentially. Gel electrophoresis of DNA with EFZ demonstrated that EFZ also has the ability to cleave supercoiled plasmid DNA. Circular dichroism study showed stabilization of the right‐handed B form of CT‐DNA. All results suggest that EFZ interacts with CT‐DNA via an intercalative mode of binding. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

9.
The interaction of four benzothiazole tethered triazole analogues (MS43, MS70, MS71, and MS78) with human serum albumin (HSA) was investigated using various spectroscopic techniques (ultraviolet–visible (UV–vis) light absorption, fluorescence, circular dichroism (CD), molecular docking and density functional theory (DFT) studies). Fluorescence quenching constants (~1012) revealed a static mode of quenching and binding constants (Kb ~104) indicating the strong affinity of these analogues for HSA. Further alteration in the secondary structure of HSA in the presence of these analogues was also confirmed by far UV–CD spectroscopy. The intensity loss in CD studied at 222 nm indicated an increase in random coil/β‐sheet conformations in the protein. Binding energy values (MS71 (?9.3 kcal mol?1), MS78 (?8.02 kcal mol?1), MS70 (?7.16 kcal mol?1) and MS43 (?6.81 kcal mol?1)) obtained from molecular docking revealed binding of these analogues with HSA. Molecular docking and DFT studies validated the experimental results, as these four analogues bind with HSA at site II through hydrogen bonding and hydrophobic interactions.  相似文献   

10.
In this study, the plant node was dried in an oven (40, 50 and 60 °C), shade and temperature-controlled microwave (40, 50 and 60 °C) methods. Statistically (p<0.05), the values closest to the color values of fresh grass were determined in an oven at 40 °C drying temperature. Effective diffusion values varied between 8.85×10−8–5.65×10−6 m2 s−1. While the activation energy was 61.28 kJ mol−1 in the oven, it was calculated as 85.24 kJ mol−1 in the temperature-controlled microwave. Drying data was best estimated in the Midilli-Küçük (R2 0.9998) model oven at 50 °C. The highest SMER value was calculated as 0.0098 kg kWh−1 in the temperature-controlled microwave drying method. The lowest SEC value in the temperature-controlled microwave was determined as 24.03 kWh kg−1. It was determined that enthalpy values varied between −2484.66/−2623.38 kJ mol−1, entropy values between −162.04/−122.65 J mol−1 and Gibbs free energy values between 453335.22–362581.40 kJ mol−1. Drying rate values were calculated in the range of 0.0127–0.9820 g moisture g dry matter−1 in the temperature-controlled microwave, 0.0003–0.0762 g dry matter−1 in the oven, and 0.001–0.0058 g moisture g dry moisture matter−1 in the shade. Phenolic content 6957.79 μg GAE g−1 fw - 48322.27 μg GAE g−1 dw, flavonoid content 3806.67 mg KE L−1 fw - 22200.00 mg KE L−1 dw and antioxidant capacity 43.35 μmol TE g−1 fw - 323.47 μmol TE g−1 dw. The highest chlorophyll values were obtained from samples dried in an oven at 40 °C. According to the findings, it is recommended to dry the knotweed (Polygonum cognatum Meissn.) plant in a temperature-controlled microwave oven at low temperatures. In this study, in terms of drying kinetics and energy parameters, a temperature-controlled microwave dryer of 60 °C is recommended, while in terms of quality characteristics, oven 40 °C and shade methods are recommended.  相似文献   

11.
BackgroundSIMR1281 is a potent anticancer lead candidate with multi- target activity against several proteins; however, its mechanism of action at the molecular level is not fully understood. Revealing the mechanism and the origin of multitarget activity is important for the rational identification and optimization of multitarget drugs.MethodsWe have used a variety of biophysical (circular dichroism, isothermal titration calorimetry, viscosity, and UV DNA melting), biochemical (topoisomerase I & II assays) and computational (molecular docking and MD simulations) methods to study the interaction of SIMR1281 with duplex DNA structures.ResultsThe biophysical results revealed that SIMR1281 binds to dsDNA via an intercalation-binding mode with an average binding constant of 3.1 × 106 M−1. This binding mode was confirmed by the topoisomerases' inhibition assays and molecular modeling simulations, which showed the intercalation of the benzopyrane moiety between DNA base pairs, while the remaining moieties (thiazole and phenyl rings) sit in the minor groove and interact with the flanking base pairs adjacent to the intercalation site.ConclusionsThe DNA binding characteristics of SIMR1281, which can disrupt/inhibit DNA function as confirmed by the topoisomerases' inhibition assays, indicate that the observed multi-target activity might originate from ligand intervention at nucleic acids level rather than due to direct interactions with multiple biological targets at the protein level.General significanceThe findings of this study could be helpful to guide future optimization of benzopyrane-based ligands for therapeutic purposes.  相似文献   

12.
《Aquatic Botany》2005,81(4):315-325
Effects of irradiance and water flow on formation and growth of filamentous and spongy thalli of Codium fragile (Suringar) Hariot growing on vinylon threads were investigated at the laboratory culture. They showed clear differences in their irradiance and water flow requirements for their formation and growth. Spongy thalli were formed from the cultured filamentous thalli only at the high water flow velocity (10 cm s−1). Number of the spongy thalli remarkably increased with increasing irradiance because those at 10, 50 and 100 μmol m−2 s−1 reached 0, 2 and 76 thalli m−1, respectively, by 10 weeks of culture. In contrast, filamentous thalli were formed from the cultured spongy thalli at 0 and 3 cm s−1, and difference in irradiance had no effect on their formation. Growth of the spongy thalli greatly accelerated under the combination of the high irradiance and high water velocity (200 μmol m−2 s−1 and 10 cm s−1) because their relative growth rate in wet weight under the condition was two–four times higher than those at the other examined irradiances and water velocities. On the other hand, difference in water velocity had no effect on growth of the filamentous thalli under flowing water, and their growth decelerated at the high irradiance (200 μmol m−2 s−1). This demonstrates that water flow is a major factor controlling the formation of the spongy and filamentous thalli. The formation and growth of the spongy thalli surely occur under the combination of the high irradiance and fast flowing water. In contrast, the formation of the filamentous thalli occurs in the calm water, and their growth is inhibited under the high irradiance.  相似文献   

13.
Rational modification of known drug candidates to design more potent ones using computational methods has found application in drug design, development, and discovery. Herein, we integrate computational and theoretical methodologies to unveil rivastigmine derivatives as dual inhibitors of acetylcholinesterase (AChE) and butyrylcholinesterase (BuChE) for Alzheimer's disease (AD) management. The investigation entails pharmacokinetics screening, density functional theory (DFT) mechanistic study, molecular docking, and molecular dynamics (MD) simulation. We designed over 20 rivastigmine substituents, subject them to some analyses, and identified RL2 with an appreciable blood-brain barrier score and no permeability glycoprotein binding. The compound shows higher acylation energy and a favored binding affinity to the cholinesterase enzymes. RL2 interacts with the AChE and BuChE active sites showing values of −41.1/−39.5 kcal mol−1 while rivastigmine binds with −32.7/−30.7 kcal mol−1 for these enzymes. The study revealed RL2 (4-fluorophenyl rivastigmine) as a potential dual inhibitor for AChE and BuChE towards Alzheimer's disorder management.  相似文献   

14.
The binding interaction of lovastatin with calf thymus DNA (ct‐DNA) was studied using UV/Vis absorption spectroscopy, fluorescence emission spectroscopy, circular dichroism (CD), viscosity measurement and molecular docking methods. The experimental results showed that there was an obvious binding interaction of lovastatin with ct‐DNA and the binding constant (Kb) was 5.60 × 103 M–1 at 298 K. In the binding process of lovastatin with ct‐DNA, the enthalpy change (ΔH0) and entropy change (ΔS0) were –24.9 kJ/mol and –12.0 J/mol/K, respectively, indicating that the main binding interaction forces were van der Waal's force and hydrogen bonding. The molecular docking results suggested that lovastatin preferred to bind on the minor groove of different B‐DNA fragments and the conformation change of lovastatin in the lovastatin–DNA complex was obviously observed, implying that the flexibility of lovastatin molecule plays an important role in the formation of the stable lovastatin–ct‐DNA complex. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

15.
A series of potent inhibitors of histone deacetylase-8 (HDAC8) is described that contains an α-amino amide zinc-binding unit and a substituted isoindolinyl capping group. The presence of a 2,4-dichlorophenyl unit located in the acetate-release cavity was shown to confer a gain of approx. 4.3 kJ mol−1 in binding energy compared to a phenyl group, and the isoindoline linker has approx. 5.8 kJ mol−1 greater binding energy than the corresponding tetrahydroisoquinoline ring system. In a series of 5-substituted isoindolin-2-yl inhibitors, a 5-acetylamino derivative was found to be more potent than the 5-unsubstituted lead HDAC8 inhibitor (increase in binding energy of 2.0 kJ mol−1, ascribed to additional binding interactions within the Nε-acetyl-l-lysine binding tunnel in HDAC8, including hydrogen bonding to Asp101. Tolerance of a 5-substituent (capping group) on the isoindoline ring has been demonstrated, and which in some cases confers improved enzyme inhibition, the HDAC8 substrate-binding region providing a platform for additional interactions.  相似文献   

16.
Binding of [3H]aflatoxin B1 to rat plasma was investigated in vivo and vn vitro. Column chromatographic and polyacrylamide gel electrophoretic analysis clearly demonstrated that aflatoxin B1 bound primarily plasma albumin. Very little binding activity was shown by other plasma proteins. Spectrofluorimetric studies were undertaken to gain some insight into the nature of the aflatoxin-albumin interaction. Quenching of the lone tryptophan fluorescence intensity upon aflatoxin binding was due, at least in part, to a ligand-induced conformational change in the albumin molecule. Aflatoxin B1 binds an apolar site with an association constant of 30 mM−1 at pH 7.4 and 20°C. Neither charcoal treatment of rat albumin nor the presence of 0.15 M NaCl had many significant effect on the interaction. The association constant was pH-dependent, increasing about 1.7-fold as the pH increased from 6.1 to 8.4. This pH dependence is ascribed to a pH-induced conformational change in the albumin molecule. Thermodynamic studies indicated that the aflatoxin-albumin interaction was exothermic (ΔH = −29.3 kJ·mol), with a ΔS value of −13.8 J·mol−1·K−1.  相似文献   

17.
Five new α-aminophosphonates are synthesized and characterized by EA, FT-IR, 1H NMR, 13C NMR, 31P NMR, ESI-MS and X-ray crystallography. The X-ray analyses reveal that the crystal structures of 1–5 are monoclinic or triclinic system with the space group P 21/c, P  1, P  1, P2(1)/c and P  1, respectively. All P atoms of 1–5 have tetrahedral geometries involving two O-ethyl groups, one Cα atom, and a double bond O atom. The binding interaction of five new α-aminophosphonate N-derivatives (1–5) with calf thymus(CT)-DNA have been investigated by UV–visible and fluorescence emission spectrometry. The apparent binding constant (Kapp) values follows the order: 1 (3.38 × 105 M−1) > 2 (3.04 × 105 M−1) > 4 (2.52 × 105 M−1) > 5 (2.32 × 105 M−1) > 3 (2.10 × 105 M−1), suggesting moderate intercalative binding mode between the compounds and DNA. In addition, fluorescence spectrometry of bovine serum albumin (BSA) with the compounds 1–5 showed that the quenching mechanism might be a static quenching procedure. For the compounds 1–5, the number of binding sites were about one for BSA and the binding constants follow the order: 1 (2.72 × 104 M−1) > 2 (2.27 × 104 M−1) > 4 (2.08 × 104 M−1) > 5 (1.79 × 104 M−1) > 3 (1.17 × 104 M−1). Moreover, the DNA cleavage abilities of 1 exhibit remarkable changes and the in vitro cytotoxicity of 1 on tumor cells lines (MCF-7, HepG2 and HT29) have been examined by MTT and shown antitumor effect on the tested cells.  相似文献   

18.
Seagrasses worldwide are highly vulnerable to, and at increasing risk from reduced light availability, and robust light thresholds are required for evaluating future impacts of changing light conditions. We tested the morphological response (shoot density and growth) of four Indo-West Pacific seagrass species (Cymodocea serrulata, Halodule uninervis, Halophila ovalis and Zostera muelleri) to six daily light levels ranging from 0 to 23 mol m−2 d−1 (0–70% surface irradiance) in cool (∼23 °C) and warm temperatures (∼28 °C) over 14 weeks. The impact of light limitation on shoot densities and growth rates was higher at warm than at cool temperatures, and for Z. muelleri and H. ovalis than for C. serrulata and H. uninervis, in terms of both the time taken for the low light treatment to take effect and the predicted time to shoot loss (e.g. 17–143 days at 0 mol m−2 d−1). Using fitted curves we estimated temperature-dependent thresholds (with estimates of uncertainty) for 50% and 80% protection of growth and shoot density, defined here as “potential light thresholds” in recognition that they were derived under experimental conditions. Potential light thresholds that maintained 50% and 80% of seagrass shoot density fell within the ranges 1.1–5.7 mol m−2 d−1 and 3.8–10.4 mol m−2 d−1, respectively, depending on temperature and species. Light thresholds calculated in separate in situ studies for two of the same species produced comparable results. We propose that the upper (rounded) values of 6 mol m−2 d−1 and 10 mol m−2 d−1 can be used as potential light thresholds for protecting 50% and 80% of shoot density for these four species over 14 weeks. As management guidelines should always be more conservative than thresholds for biological declines, we used error estimates to provide a quantitative method for converting potential light thresholds into guidelines that satisfy this criterion. The present study demonstrates a new approach to deriving potential light thresholds for acute impacts, describes how they can be applied in management guidelines and quantifies the timescales of seagrass decline in response to light limitation. This method can be used to further quantify cumulative impacts on potential light thresholds.  相似文献   

19.
Insights into binding efficacy and thermodynamic aspects of small molecules are important for rational drug designing and development. Here, the interaction of Harmane (Har), a very important bioactive indole alkaloid, with AT and GC hairpin duplex−DNAs has been reported using various biophysical tools. Detailed molecular mechanism with special emphasis on binding nature, base specificity, and thermodynamics have been elucidated via probing nucleic acids with varying base compositions. Har bound to both the DNA strands exhibited hypochromic effect in absorbance whereas bathochromic and hypochromic effects in fluorescence spectra. The binding constants estimated were in the order of 105 M−1 (higher for GC sequence compared with AT) with 1:1 stoichiometry. Noncooperative binding mode has been observed via intercalation in both the cases. The thermodynamic profile was obtained from temperature-dependent fluorescence experiments. Both Har–AT and Har–GC complexations were exothermic in nature associated with positive entropy and negative enthalpy changes. Salt-dependent studies revealed that the binding interaction was governed by nonpolyelectrolytic and hydrophobic interaction forces. The ligand-induced structural perturbation of the DNA structures was evident from the circular dichroism data. Molecular modelling data indicated towards the involvement of hydrophobic forces and hydrogen bonding.  相似文献   

20.
Empagliflozin (EMP) is an oral antihyperglycemic agent for type 2 diabetic patients. The molecular binding of EMP to bovine serum albumin (BSA) was elucidated by a combined experimental/computational approach to fulfil the pharmacokinetics and pharmacodynamics gaps of the cited drug for further development. Fluorescence, synchronous, and three-dimensional fluorescence spectroscopy verified that EMP quenched BSA native fluorescence through a dual static/dynamic mechanism that was further supported by Fӧrster resonance energy transfer and ultraviolet absorption spectroscopy. Fourier transform infrared spectroscopy revealed the conformational variations in BSA secondary structure induced by EMP. Thermodynamic properties of the BSA–EMP complex were also investigated, and the hydrophobic interactions' role in the binding process was demonstrated by the computed enthalpy (ΔH = 6.558 kJ mol−1) and entropy (ΔS = 69.333 J mol−1 K−1). Gibbs free energy (ΔG) values were negative at three distinct temperatures, illuminating the spontaneity of this interaction. In addition, molecular docking studies depicted the optimal fitting of EMP to BSA on Site I (sub-domain IIA) through three hydrogen bonds. Additionally, and based on the quenching effect of EMP on BSA fluorescence, this study suggests a simple validated spectrofluorometric method for the quantitation of the studied drug in bulk form and human plasma samples with reasonable recoveries (96.99–103.10%).  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