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 26 毫秒
1.
Sabrina Duschner 《Polymer》2006,47(21):7391-7396
Quaternized polymer combs based on poly(2-vinylpyridine-macromonomers) and the surfactant sodium dodecylsulfate are employed in the synthesis of a novel cylindrical polyelectrolyte-comb-surfactant complex (PECSC). The complex formed has 1:1 stoichiometry with respect to the ratio of dodecylsulfate to pyridinium units. It is soluble in organic solvents such as 2-butanol or chloroform. Characterization of single particle properties of the complex in organic solution is possible and yields a radius of gyration of 〈Rgz = 78.4 nm, a hydrodynamic radius of 〈1/Rhz−1 = 51.4 nm and a cross-sectional radius of Rg,cross = 3.9 nm in chloroform. The characteristic ratio γ = 〈Rgz/〈1/Rhz−1 decreases from γ = 1.73 for the original quaternized polymer comb, indicating the semi-flexible, cylindrical nature of the macromolecules in aqueous solution, to γ = 1.53 for the complex in chloroform. The effect of the main-chain stretching accompanied by the increase of the volume of the comb by introduction of the surfactant is smaller as compared to the electrostatic interactions in the parent comb. This is also reflected in the persistence length lp, which is determined by SANS, and found to be 21.3 nm for the complex and 24.4 nm for the polyelectrolyte comb. In addition, AFM investigations of the polymers adsorbed onto mica showed a 2D-equilibrium adsorption for the complex and a kinetically dominated adsorption process in case of the polyelectrolyte comb.  相似文献   

2.
Ashok K. Das 《Polymer》2010,51(10):2244-30
Translocation of polymer chains under the application of an external force has been studied through coarse-grained Monte Carlo simulations. The chains are pulled through a nanotube of finite length and diameter and their translocation times measured. The average translocation time, τ follows a scaling relation involving the chain length, N and applied force, F as, τ ∼ Nν′F−μ, where ν′ and μ are two different exponents (ν′ = 0.674, and μ = 0.95 ± 0.05). The scaling law is closely similar to the nanopore translocation scaling law reported by Milchev et al. [Ann N Y Acad Sci 2009;1161:95]. Characteristic signatures of the chain escape time have been exhibited by the square of end-to-end distance R2, axial radius of gyration Rg−x and other constituent properties. The behavior of the linear polymers under the application of a pulling force has been exploited to gain insights into the ultrafiltration process of unentangled polymers in dilute solution. The generic pulling force-translocation time (F, τ) data obtained through simulation can be matched reasonably well with the hydrodynamic force-critical macroscopic flow time (fh, Qc−1) data and also with the hydrodynamic force-reduced critical microscopic flow time (fh, qc−1) data obtained in the ultrafiltration experiment on long linear polystyrene chains in cyclohexane, as recently reported by Ge et al. [Macromolecules 2009;42:4400] The simulation technique reported here may be extended to study biomolecular transports occurring in long protein channels, as studied experimentally through current-time or voltage-time traces.  相似文献   

3.
Partially crystalline bisphenol A polycarbonate (BPAPC) nanofibers were successfully fabricated using a combination of a centrifugal field (1800 rpm) and an electrostatic field (25 kV). The BPAPC solution properties are key factors for adequately electrospinning the partially crystalline BPAPC nanofibers. The correlation times (τc) of methyl (τc = 9.3 ns) and of benzene-ring (τc = 15.3 and 15.8 ns) motions in the 14 wt.% BPAPC/THF solution were longer than in CH2Cl2 and CHCl3, as determined by NMR. The distribution-peak maximum of the hydrodynamic radius of BPAPC in the 14 wt.% THF solution (Rh = 15 Å) was higher than in CH2Cl2 (Rh = 9.2 Å) and CHCl3 (Rh = 7.9 Å), as evidenced by DLS data. We conclude that the BPAPC assumed a denser, more worm-like chain conformation in THF solvation.  相似文献   

4.
The propagation and termination rate coefficients for bulk polymerization of the butyl acrylate dimer (BA dimer) are determined by pulsed laser techniques. The rate coefficient for propagation, kp, is deduced for temperatures from 20 to 90 °C via the pulsed laser polymerization-size exclusion chromatography (PLP-SEC) method at pulse repetition rates between 1 and 10 Hz. The Arrhenius parameters were found to be: EA(kp) = (34.2 ± 1.0) kJ mol−1 and A(kp)/L mol−1 s−1 = (1.08 ± 0.49) × 107 L mol−1 s−1. The termination rate coefficient, kt, has been measured via SP-PLP-ESR, single pulse-pulsed laser polymerization in conjunction with time-resolved electron spin resonance detection of radical concentration. The resulting Arrhenius parameters as deduced from the temperature range −15 to +30 °C are: EA(〈kt〉) = (22.8 ± 3.7) kJ mol−1 and log(A/L mol−1 s−1) = 10.6 ± 1. The chain-length dependence of kt was studied at 30 °C. For short chains a significant dependence was found which may be represented by an exponent α = 0.79 in the power-law expression kt(i) = kt0i−α.  相似文献   

5.
This study shows that the degradation mechanism and kinetics of monodisperse oligo(lactic acid)s esterified with N-(2-hydroxypropyl)methacrylamide (HPMAm) are strongly influenced by the nature of the chain end. Oligomers with free hydroxyl chain ends degraded predominantly by chain end scission via a backbiting mechanism with a pseudo first-order rate constant kbb=2.7 h−1 in aqueous buffer (37 °C, pH 7.2). Once the hydroxyl groups were protected by acetylation, random chain scission became the rate limiting step with kr=0.022 h−1 under the same conditions. Using these rate constants, the theoretical time-resolved degradation profile was calculated for every (intermediate) degradation product and corresponded very well with the experimental results. The rate of formation of HPMAm was independent of the chain length for the acetylated oligomers, while the hydroxyl terminated oligomers with an even number of lactic acid units formed HPMAm more rapidly than oligomers with an odd number of units. The possibility to fine-tune the degradation rate is relevant when applied as e.g. hydrogels for controlled release or tissue engineering materials.  相似文献   

6.
Shane A. Seabrook 《Polymer》2005,46(23):9562-9573
The kinetics of acrylamide (AAm) free radical polymerization at low conversion of monomer to polymer in the aqueous phase was investigated at 50 °C using γ-radiolysis relaxation, which is sensitive to radical-loss processes. The values of the termination rate coefficients for AAm ranged from 8×106 to 3×107 M−1 s−1 as the weight fraction of polymer ranged from 0.002 to 0.0035, which is significantly lower than the low-conversion values for monomers such as styrene (2×108 M−1 s−1) and methyl methacrylate (4×107 M−1 s−1) in organic media. These can be quantitatively explained by applying a chain-length-dependent model of free-radical polymerization kinetics [Russell GT, Gilbert RG, Napper DH. Macromolecules 1992;25:2459. [19]] in which termination kinetics are expressed in terms of a diffusion-controlled encounter of radicals which ultimately yields an expression for the chain-length-averaged termination rate coefficient, 〈kt〉. The lower 〈kt〉 for AAm arises due to a combination of the high kp value, promoting rapid formation of slower terminating long chains, and the slow diffusion of short propagating chains, relative to other common monomers. The chain transfer to monomer constant for AAm in water at 50 °C, CM, was estimated using the chain-length-distribution method with correction for band-broadening [Castro JV, van Berkel KY, Russell GT, Gilbert RG. Aust J Chem 2005;58:178. [21]] and found to be 1.2×10−4 (±10%). The diffusion characteristics for AAm were adapted from those obtained for a similar aqueous system (hydroxyethyl methacrylate) together with a 0.5 exponent for the power law dependence on penetrant degree of polymerization at zero weight fraction polymer. This provides an adequate fit to the 〈kt〉 data. This is the first application of the chain-length-dependent model to describe experimental termination rate coefficients for an aqueous system at low conversion to polymer. The result that the experimental termination rate coefficients can be reproduced with an a priori model with physically reasonable parameters supports the physical assumptions underlying that model.  相似文献   

7.
Elanio A. Medeiros 《Fuel》2011,90(4):1696-1699
The rate constants for the quenching of biacetyl phosphorescence by a series of conjugated dienes were measured. 1,3-cyclohexadiene (kqP = 2.94 × 109 s−1 mol−1 L), 2,5-dimethyl-2,4-hexadiene (kqP = 1.91 × 109 s−1 mol−1 L), 2,4-dimethyl-1,3-pentadiene (kqP = 1.78 × 108 s−1 mol−1 L), 3-methyl-1,3-pentadiene (kqP = 1.22 × 108 s−1 mol−1 L), 2,4-hexadiene (kqP = 1.35 × 108 s−1 mol−1 L) and trans-2-methyl-1,3-pentadiene (kqP = 3.84 × 108 s−1 mol−1 L). Cyclooctene also quenched biacetyl phosphorescence but with a lower rate (kqP = 1.97 × 107 s−1 mol−1 L). Quenching was not observed with 1-methylnaphthalene. Since conjugated dienes quench biacetyl phosphorescence preferentially, this method was studied using gasoline samples with known diene composition. A good correlation was found between the rate of quenching of biacetyl by the gasoline samples and the quantity of conjugated dienes present.  相似文献   

8.
This work evaluates the volumetric mass transfer coefficient (kLa), the gas hold-up (?) and the mixing time (tm) as a function of superficial gas velocity (UG) in a flat-panel photobioreactor (PBR) with high light path. CO2 utilization efficiency and volumetric power consumption (P/V) were also evaluated. A 50 L working volume photobioreactor was developed, 0.67 m in length, 0.57 m in height and 0.15 m in width (light path). The height-width ratio was 3.8, which is lower than reported in most PBRs. Initially, experiments were performed with air and tap water (biphasic system) and, subsequently, using a Spirulina sp. culture (triphasic system: air, culture medium, cells). Minimum and maximum superficial gas velocity values were 5 × 10−5 and 8.4 × 10−3 m s−1, respectively. Maximum values for kLa and ? were 20.34 h−1 (0.0057 s−1) and 0.033 in the biphasic system, and 31.27 h−1 (0.0087 s−1) and 0.065 in the triphasic system. CO2 utilization efficiency was 30.57%. Results indicate that the hydrodynamic and mass transfer characteristics of this photobioreactor are more efficient than those reported elsewhere for tubular and other flat-plate PBRs, which opens the possibility of using PBRs with higher light paths than yet proposed.  相似文献   

9.
An alumina precursor was prepared by the aluminium sulphate (0.20 M) and excess urea reaction in boiling aqueous solution. The precursor was calcined at 900 °C for 2 h and then δ-Al2O3 powder having volumetric agglomeration degree of 80% was obtained. Cylindrical compacts having diameter of 14 mm were prepared under 32 MPa by axial pressing using oleic acid as binder. Each compact was fired isothermally at various temperatures between 950 and 1400 °C. The firing time was changed from zero to 2 h. The fired compacts were examined by scanning electron microscopy (SEM) and nitrogen adsorption techniques. The specific surface areas (S/m2 g− 1) of the samples were calculated using the Brunauer, Emmett, and Teller (BET) procedure. The rate constant (k) and mechanism-characteristic parameter (n) were obtained for different temperatures between 950 °C and 1150 °C from the application of the neck-growth sintering rate (NGRM) model on the surface area reduction data. An Arrhenius equation and the parameter n for the sintering were found in the forms of k = (7.648 × 106 h− 1) exp (− 186,234 J mol− 1 / RT) and n = 4.0 × 10− 7 T3-1.7 × 10− 3 T2 + 2.3 T − 1030.8 respectively. The parameter n changes in the interval 0.61 <  n < 1.34 with rising temperature having maximum at about 1025 °C. Based on the SEM images and NGRM data, the intra-particle sintering was discussed.  相似文献   

10.
The standard rate constant of a simple electrode reaction Ox + ne ↔ Red, in which both Ox and Red are solution soluble, can be determined by the variation of frequency in the square-wave voltammetry with inverted scan direction: log ks = log f01/2 + log D1/2 − 0.60 ± 0.01. In this equation log f0 is the abscissa of the intersection of straight lines Ep,2 = a log f + b and Ep,2 = E0, where Ep,2 is the potential of the second peak of the net response, E0 is the standard potential, a = 2.3RT/2(1 − α)nF, b = E0 − 2a log ks + a log D − 0.0353/(1 − α)n and D is a common diffusion coefficient.  相似文献   

11.
Cheng Yang  Bo Meng  Mingqing Chen  Zhongbin Ni 《Polymer》2006,47(23):8044-8052
The dynamic behaviors of potato amylopectin and waxy corn amylopectin in semidilute solution were investigated by laser light scattering and viscometer. For potato amylopectin with relatively smaller molecular weight, only pure diffusion motion of amylopectin was found by LLS in dilute regime. When the concentration was above the critical overlapping concentration (C), three relaxation modes were found. The line-width of fast mode (Γf) had a q2 dependence, where q is the scattering vector, and the correlative length (〈ξh〉) could be scaled to concentration (C) as 〈ξh〉 ∼ C−0.79±0.1 when C > 2%. This mode was attributed to the cooperative relaxation motion of the “blobs” in the transient network. The line-width of slow relaxation mode (Γs) could be scaled with q as Γs∼qαs, αs varying from 2.0 to 2.66 as the concentration increased. The relaxation time of slow relaxation mode (τs) had a C1.8±0.1 dependence. This mode was originated from the association of the amylopectin. The medium mode was found when C > 4%. The line-width of medium relaxation mode (Γm) could be scaled to q as Γm∼qαm, αm varying from 2.7 to 2.5 with the increasing concentration. The relaxation time of medium relaxation mode (τm) had C0.7±0.1 dependence. The relative intensity contribution of the medium relaxation mode decreased with a rise in the concentration. This mode was attributed to the thermally agitated density fluctuation in semidilute solution induced by heterogeneities of the transient network. For waxy corn amylopectin with relatively huge molecular weight (∼108 g/mol), only the internal motion of the single amylopectin molecule was found in dilute regime when qRg ≥ 2, where Rg is the gyration radius of amylopectin. It was also found that there were three relaxation modes in semidilute solution of waxy corn amylopectin. The fast relaxation mode was found to be caused first by the internal motion of the single amylopectin molecule, and then, with the increasing concentration, by the cooperative motion of the transient network. The medium and slow relaxations for waxy corn amylopectin have the same physical origin as those for potato amylopectin. However, the C dependence and the q dependence of the medium and slow relaxation times for waxy corn amylopectin were different from those for potato amylopectin. This was attributed to the strong dynamic coupling effect in semidilute solution of the waxy corn amylopectin. The concentration dependence of the viscosity of amylopectin in semidilute solution indicated that the topological entanglement of amylopectin was weak due to the highly branching.  相似文献   

12.
J.I. Cail  I.M. Ward 《Polymer》2007,48(5):1367-1378
The Monte-Carlo (MC) approach of Paper I is developed to predict the birefringence of PET. An extension of the modelling of polarisability is used that accounts for chain flexibility and for structural units containing several types of bonds. The rotational-isomeric-state (RIS) model for PET chains in melts is employed to calculate the polarisability of the terephthaloyl segment and of each of the 27 possible conformations of the five skeletal bonds of the glycol segment. These polarisabilities then enable the birefringent properties of drawn PET melts to be predicted. A method of pre-averaging the individual glycol polarisabilities that greatly reduces the length of the calculation is shown to be valid.It is found that shorter PET chains produce higher values of birefringence (Δñ) for a given deformation. This trend is due to greater proportions of the chains reaching higher conformational extensions and therefore becoming more oriented. In disagreement with Kuhn and Grün theory, Δñ is not linearly related to λ2 − λ−1. This non-linear behaviour is related to the non-linear behaviour of the orientation functions of the terephthaloyl and glycol segments, 〈P2(cos ζter)〉 and 〈P2(cos ζgly)〉, with λ2 − λ−1. The non-linear behaviour of 〈P2(cos ζter)〉 with λ2 − λ−1 was confirmed experimentally in Paper I, where the measured values of 〈P2(cos ζter)〉 were found to be closely predicted by the present MC modelling.In the present paper, the predicted values of Δñ, calculated according to various published values of bond polarisabilities are presented and discussed. They will be used in Paper III to model the measured birefringence of drawn PET.  相似文献   

13.
Hydrogen sulfide (H2S) is currently removed from gaseous effluents by chemical scrubbing using water. Chlorine is a top-grade oxidant, reacting with H2S with a fast kinetic rate and enhancing its mass transfer rate. To design, optimize and scale-up scrubbers, knowledge of the reaction kinetics and mechanism is requested. This study investigates the H2S oxidation rate by reactive absorption in a mechanically agitated gas–liquid reactor. Mass transfer (gas and liquid sides mass transfer coefficients) and hydrodynamic (interfacial area) performances of the gas–liquid reactor were measured using appropriated physical or chemical absorption methods. The accuracy of these parameters was checked by modeling the H2S absorption in water without oxidant. A sensitivity analysis confirmed the robustness of the model. Finally, reactive absorption of H2S in chlorine solution for acidic or circumneutral pH allowed to investigate the kinetics of reaction. The overall oxidation mechanism could be described assuming that H2S is oxidized irreversibly by both hypochlorite anion ClO (k = 6.75 × 106 L mol−1 s−1) and hypochlorous acid ClOH (k = 1.62 × 105 L mol−1 s−1).  相似文献   

14.
The structure of the free volume and its temperature dependence in poly(tetrafluoroethylene) (PTFE) and of its copolymer with perfluoro(propyl vinyl ether) (PFA) was studied by pressure-volume-temperature (PVT) experiments (T=27-380 °C, P=0.1-200 MPa) and positron annihilation lifetime spectroscopy (PALS, T=−173-250 °C, P=10−5 Pa). From the analysis of these experiments we conclude on the volumetric properties of the mobile (MAF) and rigid amorphous fractions (RAF) in these semicrystalline polymers. The specific volumes of the MAF and RAF, VMAF and VRAF, were estimated assuming that VMAF agrees with the specific volume of the melt extrapolated down to lower temperatures using the Simha-Somcynsky equation of state (S-S eos). VRAF was then estimated from the specific volume of the entire amorphous phase, Va, and the known VMAF. The specific free volume Vf=VaVocc was also estimated from Va using the S-S eos hole fraction h, Vocc=(1−h)Va. From the analysis of PALS data with the routine LT9.0 the mean volume, 〈vh〉, and the width, σh, of the local free volume size distribution (holes of subnanometre size) were obtained. A comparison of 〈vh〉 with Vf delivered the hole density Nh. The volume parameters show that the RAF which is formed during crystallisation from the melt has a distinctly smaller specific free and total volume than the MAF. During cooling the contraction of the RAF slows down and finally, below room temperature, the RAF possesses a larger free volume than the MAF shows. Obviously, the restriction of the segmental mobility in the RAF by the crystals limits at high temperatures the free volume expansion and at low temperatures dense packing of the polymer chains. These conclusions from the analysis of the specific volume are confirmed by PALS experiments.  相似文献   

15.
This paper reports results of molecular dynamics simulations on the electrochemical reduction of the tert-butyl bromide molecule according to the scheme: (CH3)3CBr + e → (CH3)3C + Br. A model Hamiltonian is proposed, that is based on the Anderson-Newns concept of the electron transfer process and on quantum calculations performed for the (CH3)3CBr molecule and the (CH3)3CBr· radical anion. The rate constants k and the transfer coefficients α obtained at different solvent frictions γx, temperatures T and overpotentials η, are presented. At η = 1.3 V, which corresponds to the experimental conditions, the α estimates obtained from the simulations with low solvent frictions are very close to the measured value of 0.2; in more viscous solvents they are slightly higher. The α coefficient is found to depend non-linearly on η. It is also shown to depend on the temperature: at lower overpotentials the transfer coefficient decreases with T: α (348 K) < α (298 K) < α (248 K), while for high η the reversed order appears. In all simulations strong saddle point avoidance was observed. The choice of the effective reaction paths is shown to be responsible for some trends found in the results.  相似文献   

16.
T. Jiang 《Electrochimica acta》2007,52(13):4487-4496
The kinetics of the oxygen reduction reaction (orr) on Cu(h k l) surfaces are investigated in perchloric acid and sulfuric acid solutions using rotating ring disk electrode (RRDCu(h k l)E). Parameters, such as reaction order, kinetic current, rate constant, Tafel slopes as well as the number of electrons transferred are determined. The variation in the activity and reaction pathway with the crystal faces in different electrolytes is related to the surface characteristics of Cu(h k l) and the structure-sensitive inhibiting effect of the adsorbed anions on their surfaces. In 0.1 M HClO4, the difference in activity is clearly observed on Cu(h k l) surfaces (Cu(1 0 0) > Cu(1 1 1) although it is relatively small). The higher activity of Cu(1 0 0) arises from its more open characteristics which may facilitate the co-adsorption of O2. On the other hand, the adsorption of oxygenated species on Cu(1 1 1) at E > −0.35 V induces a 2 e pathway; while a 4 e reduction is observed on Cu(1 0 0) in the entire potential region (−0.70 V < E < −0.10 V). In 0.5 M H2SO4, the sequence in activity between Cu(1 1 1) and Cu(1 0 0) varies with the potentials, i.e., Cu(1 0 0) is initially more active than Cu(1 1 1) at −0.35 V < E < −0.15 V, however, the reversal in the activity between Cu(1 1 1) and Cu(1 0 0) is observed at more negative potentials (−0.45 V < E < −0.35 V). The desorption of strongly adsorbed (bi)sulfate anions on Cu(1 1 1) induces the 2 e reduction via peroxide formation, however, a 4 e reduction is dominant on the Cu(1 0 0) surfaces. The major effect of (bi)sulfate anions and oxygenated species on the orr kinetics and reaction pathway on Cu(h k l) surfaces is the blocking of active copper sites for the adsorption of O2 molecules.  相似文献   

17.
Fang Ye  Lishi Wang 《Electrochimica acta》2008,53(12):4156-4160
5-[o-(4-Bromine amyloxy)phenyl]-10,15,20-triphenylporphrin (o-BrPETPP) was electropolymerized on a glassy carbon electrode (GCE), and the electrocatalytic properties of the prepared film electrode response to dopamine (DA) oxidation were investigated. A stable o-BrPETPP film was formed on the GCE under ultrasonic irradiation through a potentiodynamic process in 0.1 M H2SO4 between −1.1 V and 2.2 V versus a saturated calomel electrode (SCE) at a scan rate of 0.1 V s−1. The film electrode showed high selectivity for DA in the presence of ascorbic acid (AA) and uric acid (UA), and a 6-fold greater sensitivity to DA than that of the bare GCE. In the 0.05 mol L−1 phosphate buffer (pH 6.0), there was a linear relationship between the oxidation current and the concentration of DA solution in the range of 5 × 10−7 mol L−1 to 3 × 10−5 mol L−1. The electrode had a detection limit of 6.0 × 10−8 mol L−1(S/N = 3) when the differential pulse voltammetric (DPV) method was used. In addition, the charge transfer rate constant k = 0.0703 cm s−1, the transfer coefficient α = 0.709, the electron number involved in the rate determining step nα = 0.952, and the diffusion coefficient Do = 3.54  10−5 cm2 s−1 were determined. The o-BrPETPP film electrode provides high stability, sensitivity, and selectivity for DA oxidation.  相似文献   

18.
Polymer translocation through a nanopore in mesoscopic simulations   总被引:1,自引:0,他引:1  
Yan-Dong He  Ze-Sheng Li 《Polymer》2007,48(12):3601-3606
Dissipative particle dynamics (DPD) simulations are carried out to study the translocation of a single polymer chain through a pore under fluid field. The influences of the field strength E, the chain length N, the solvent quality αsp, and the pore size h on the translocation time are evaluated. The translocation time τ, which is defined as the time that the chain moves through the pore completely in the direction of the driving force, scales with the field strength E as τ ∼ E−0.48±0.01. We find that the translocation time is proportional to the chain length, which is in agreement with the experimental results and theoretical predictions. Tracing the variation of the square radius of gyration, , and the polymer configuration during translocation, we observe that the chain is elongated when it is passing through the pore, which manifests that the chain is not in equilibrium during the translocation process. We also find that the worse the solvent quality is, the less time it will take to translocate, no matter what the size of the pore is. If the size of the pore is enlarged, the translocation time will be shorter. The information we gain from this study may benefit to the DNA sequencing.  相似文献   

19.
The complex of rutin-Cu (C81H86Cu2O48, abbreviated by Cu2R3, R = rutin) was synthesized and characterized by elemental analysis and IR spectra. Cyclic voltammetry (CV) and fluorescence spectroscopy were used to investigate the interaction of Cu2R3 with salmon sperm DNA. It was revealed that Cu2R3 could interact with double-stranded DNA (dsDNA) by a major intercalation role. Using Cu2R3 as a novel electroactive indicator, an electrochemical DNA biosensor for the detection of specific DNA fragment was developed and its selectivity for the recognition with different target DNA was assessed by differential pulse voltammetry (DPV). The target DNA related to coliform virus gene could be quantified ranged from 1.62 × 10−8 mol L−1 to 8.10 × 10−7 mol L−1 with a good linearity (r = 0.9989) and a detection limit of 2.3 × 10−9 mol L−1 (3σ, n = 7) was achieved by the constructed electrochemical DNA biosensor.  相似文献   

20.
The effect of polydispersity on dilute solution properties and microphase separation of polydisperse high-molecular-weight (Mw > 105 g mol−1) polystyrene-block-poly(styrene-co-acrylonitrile) diblock copolymers, PS-block-P(S-co-AN), was studied in this work. For experiments, a series of diblock copolymers with variable weight fractions of acrylonitrile units (wAN = 0.08-0.29) and length of block P(S-co-AN) was synthesized using nitroxide-mediated radical polymerization (NMP) technique, namely, by chain extension of nitroxide-terminated polystyrene (PS-TEMPO). According to light scattering and viscometry measurements in dilute tetrahydrofuran (THF) solutions the studied diblock copolymers assumed random coil conformation with the values of characteristic structure factor Rg/Rh = 1.50-1.76. It was found that polydisperse diblock copolymers being in strong segregation limit (SSL) self-assembled into microphase-separated ordered morphologies at ordinary temperature. The long periods of lamellar microdomains were larger compared to theoretical predictions for hypothetical monodisperse diblock copolymers. It was demonstrated by means of SAXS and TEM that a transition from a lamellar (LAM) to irregular face-centered-cubic (FCC) morphology occurred with increasing volume fraction of P(S-co-AN) block.  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