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.
It is shown that the Mott fragmentation model and the Hirsch formula for the shaped charge jet break-up time complete one another to yield a unified model. This model is consistant with the microscopic picture of internal flow splitting in the metal, into shear bands undergoing a strain larger than the average and the metal between them which undergoes a strain which is smaller than the average. This splitting reduces the deformation energy dissipated by the metal while elongating. It corresponds to the splitting of the stress versus strain characteristics of the metal into the isothermic and adiabatic curves as measured by Johnson and Cook. The parameter Vplwhich describes the break-up process according to the Hirsch model, is identified with the expression (dσm/ϱ)1/2 where dσm is the difference in yield strength between the isothermic and adiabatic curves where the adiabatic characteristic becomes a maximum. The comparison with available experimental measurements of this prediction shows a very encouraging agreement. The weak dependence of Vpl on the strain rate on the one hand and the observed increase of the jet break-up time in slowly elongating jets on the other hand are also explained using energy considerations. The recent recovery of jet particles by Zernow yields supportive evidence of the proposed theory which may be extended to all semi steady metal plastic flow phenomena. The final break-up, according to this model is due to the separation between the two sides of shear bands where the slide motion went on during the whole deformation process.  相似文献   

2.
The gas holdup and gas entrainment of a plunging liquid jet with a gas entrainment guide in an air-water system was investigated. The measurement of the gas holdup was performed using an over-flow method. The turbulent jet velocity calculated on an inside nozzle diameter in the range from 4.4-26.5 m/s for this system has been used in our correlations. The gas holdup has been well correlated in terms of 1/H(v02 + 2gH1), H1 d0 and the gas entrainment in terms of 1/Hw(v02 + 2gH1), H1, d0. The jet power requirement was also obtained from experimental data.  相似文献   

3.
Desulfurization of JP-5 jet fuel (1172 ppmw S) was investigated by π-complexation adsorption with AgNO3 supported on mesoporous silica SBA-15 and MCM-41. The average pore sizes of AgNO3/SBA-15 and AgNO3/MCM-41 were 48.8 and 19.1 Å, respectively. The results of JP-5 desulfurization showed that significant sulfur breakthrough occurred at ∼10.0 and ∼15.0 mL/g by AgNO3/SBA-15 and AgNO3/MCM-41, respectively, at a space velocity of 1.25 h−1. The spent AgNO3/MCM-41 was regenerated by a simple process (heating in air at 200 °C) and ∼50% of the sulfur capacity was recovered after the first cycle. Molecular orbital calculations show that Cu+ (as that in CuY zeolite) formed stronger π-complexation bonding with the thiophenic compounds than Ag+ (in AgNO3), as evidenced by experimental heats of adsorption. However, pore diffusion limitation of the large sulfur molecules (alkylated benzothiophenes) became an important factor for desulfurization of high sulfur jet fuels such that the AgNO3-supported mesoporous sorbents yielded substantially better results than Cu(I)Y, although Cu(I)Y was better for a model fuel that contained only small sulfur molecules. Among all sorbents that have been investigated, the AgNO3/MCM-41 sorbent showed the best desulfurization performance for high sulfur jet fuels.  相似文献   

4.
In a Confined Plunging Liquid Jet Contactor (CPLJC) a jet of liquid is introduced into an enclosed cylindrical column (downcomer) that generates fine gas bubbles that are contacted with the bulk liquid flow. The region where the liquid jet impinges the receiving liquid and expands to the wall of the downcomer is called the Mixing Zone (MZ). In the MZ, the energy of the liquid jet is dissipated by the breakup of the entrained gas into fine bubbles, and the intense recirculation of the two-phase mixture. The study presented here was undertaken to quantify the ozone-water mass transfer performance of the MZ through the determination of the volumetric mass transfer coefficient, kLa (s?1), and to produce a model for predicting kLa based on the specific energy dissipation rate. It was found experimentally that kLa in the MZ increased with increasing superficial gas velocity. A maximum experimental kLa value of 0.84 s?1 was achieved which compares well to other contactors used in water treatment. Such a large kLa value combined with the small volume of the reactor, favorable energy requirements and safety features of the system, suggests that the CPLJC provides an attractive alternative to conventional ozone contactors. The relatively large mass transfer rates were found to be a function of the high gas holdup and fine bubble size generated in the MZ, which results in an almost froth-like consistency. A model based on the specific energy dissipation rate of the water jet, E (kg · m?1· s?3), and MZ bubble size was used to predict kLa in the MZ. Using E, the number average bubble size was predicted which was then used to calculate the liquid phase mass transfer coefficient kL. The bubble size was also used with the predicted mixing zone gas holdup to calculate the specific interfacial area, a (m?1), which was then combined with kL to determine a predicted value of kLa. The average deviation between experimental and predicted kLa was 6.2%.  相似文献   

5.
A jet-impactor assisted-fluidized bed (JIAFB) for continuous de-agglomeration of nanopowder agglomerates was presented in previous work. Therein, a jet caused agglomerates to impinge up an impactor, where they would break. However, efficient impactor positioning will be dictated by particle momentum: the product of solid concentration and velocity must be highest. Herein, the variation of solid hold-up was measured in a fluidized bed of Fe2O3 nanoparticles using gamma-ray densitometry. Behaviour was compared under minimum fluidization and bubbling regimes, over a wide range of jet velocities (0–200 m s−1). A new line-decomposition approach allowed mapping local solid distribution across seven axial and five radial positions, tangibly demonstrating how increasing the gas velocity enhanced the fluidization quality by increasing axial solid diffusivity. Conversely, increasing jet velocity locally decreased solid hold-up in the jet-affected zone, and brought about inhomogeneities. With this information in hand, jet-to-impactor distance was optimized and validated experimentally.  相似文献   

6.
The performance of jet fuel depends on the density (ρ), condensed phase heat of formation (▵fH°(c)), and specific impulse (ISP). Exo‐tricyclo[5.2.1.0(2,6)]decane (C10H16) or JP‐10 is now used as a suitable synthetic liquid jet fuel because it has the approximated values of ρ=1.1 g cm−3 and ▵fH°(c)=− 123 kJ mol−1 and a broad range between the melting and boiling points, i.e. TbpTmp=196.2 K. This work introduces a suitable pathway for calculation of the values of ρ, ▵fH°(c), and ISP of 13 well‐known isomers of JP‐10 and a series of saturated polycyclic hydrocarbons with general formula of CnHn (5≤n≤12) in order to specify high performance jet fuels. Although 13 compounds have larger values of ISP*ρ than JP‐10, only two compounds, tetraspiro[2.0.0.0.2.1.1.1]undecane and tetracyclo[3.2.0.0(2,7).0(4,6)]heptane, are suitable as jet fuels.  相似文献   

7.
The homogeneous particle nucleation in a free turbulent jet is simulated at Reynolds numbers of 3400 and 34,000 by using large eddy simulation (LES) coupled with our newly developed equivalent mean nucleation method (EMNM) for the present study. The simulated velocity field, passive scalar concentration, and nucleated particle concentration are compared with various available measurements in the literature. Both the simulated mean velocity and mean scalar concentration show self-similarity properties and agree well with the experimental measurements quantitatively in the literature. Using the EMNM, it is found that the nucleation primarily occurs in a very narrow region around the jet centerline. It is also found that the position of maximum nucleation is a function of the vapor temperature and concentration at the jet exit and its surrounding background (i.e., T 0, T , c 0, and c ), and the function varies monotonically with respect to these four parameters. Present numerical simulation and analysis also show that the group, (where is the mean particle number concentration, u 0 is the jet velocity, and d is the jet exit diameter), is constant along the axial direction of the jet, which was first observed in the experiment of Lesniewski, T. K., and Friedlander, S. K. (1998). Particle Nucleation and Growth in a Free Turbulent Jet. Proc. R. Soc. Lond. A, 454:2477–2504. However, a different physical explanation of the fact is provided. In this study, it is found that the maximum nucleation occurs around 20d away from the jet exit, other than in the shear layer (less than 5d), which was supposed by Lesniewski, T. K., and Friedlander, S. K. (1998). On the basis of the additional simulated results, more comprehensive and reasonable explanations of their measurement data are provided. In this study, an efficient and rigorous numerical method is presented to simulate the homogeneous particle nucleation in a free turbulent jet, and this has provided a deeper and thorough understanding of the process.  相似文献   

8.
Munki Kim  Youngbin Yoon 《Fuel》2011,90(8):2624-2629
The effect of fuel composition on flame length was studied in a non-premixed turbulent diluted hydrogen jet with coaxial air. Because coaxial air entrained in a fuel stream enhances the mixing rate of fuel and air, it substantially reduces flame length. The observed flame length was expressed as a function of the ratio of coaxial air to fuel jet velocity and compared with a theoretical prediction based on the velocity ratio. Four cases of fuel mixed by volume were determined: 100% H2, 80% H2/20% N2, 80% H2/20% CO2, and 80% H2/20% CH4. In addition, fuel jet air velocity and coaxial air velocity were varied in an attached flame region as uF = 86-309 m/s and uA = 7-14 m/s. In this study, we derived a scaling correlation for predicting the flame length in a simple jet with coaxial air using the effective jet diameter in a near-field concept. The experimental results showed that the visible flame length was in good relation to the theoretical prediction. The scaling analysis is also valid for diluted hydrogen jet flames with varied fuel composition, which affects flame length by varying the density of the fuel.  相似文献   

9.
Poly(ethylene terephthalate) (PET) nanofibers were prepared by irradiating a PET fiber with radiation from a carbon dioxide (CO2) laser while drawing it at supersonic velocities. A supersonic jet was generated by blowing air into a vacuum chamber through the fiber injection orifice. The flow velocity from the orifice was estimated by computer simulation; the fastest flow velocity was calculated to be 401 m s−1 at a chamber pressure of 6 kPa. A nanofiber obtained using a laser power of 8 W and a chamber pressure of 6 kPa had an average diameter of 193 nm and a draw ratio of about 900,000. This technique is a novel method for producing nanofibers.  相似文献   

10.
This paper describes a novel system of a water‐sparged aerocyclone (WSA) for fine particles (FP) removal in the air stream from a coal‐fired power plant. The effects of operating parameters, including inlet FP concentration (C 0(FP)), water jet velocity (UL), and air inlet velocity (UG) on the removal efficiency of FP (REFP) were investigated. The change of morphology, particle size distribution, and content of heavy metals of the FP samples before and after removal in the WSA were also compared. The results show that the RE of the FP and PM2.5 can reach as high as 99.36 and 99 %, respectively, under optimal conditions. A regression removal model of REFP = 85.08 × ReG0.0187 × ReL0.0037 × (C0(FP)/ρFP)0.0048 was proposed to predict REFP, and the calculated REFP by the model is in agreement with experimental data with deviations of ±0.5 %. This new technology provides an alternative approach for FP removal from exhaust gas and exhibits significant potential for industrial application.  相似文献   

11.
Chi Wang  Chia-Hung Hsu  I.-Hwe Hwang 《Polymer》2008,49(19):4188-4195
Using chloroform/dimethylformamide (CF/DMF) co-solvent, electrospinning of poly[(R)-3-hydroxybutyrate] (PHB) solutions was carried out at ambient temperature. The effects of the applied voltage (V), flow-rate (Q), and solution viscoelastic properties on the Taylor cone, electrified jet, and fiber morphology were investigated. In addition, the electric field developed by the needle-plate electrode configuration was calculated using a finite element analysis to reveal the tip-to-collector (H) effect. Among the processing parameters (V, Q and H), it was found that Q played a key role in determining the jet diameter (dj) and electrospun fiber diameter (df), and scaling laws existed between them, i.e., dj-Q0.61 and df-Q0.33. The diameter reduction ratios of Do/dj (Do is the needle diameter) and dj/df were measured as 50-120 and 5-10, respectively; it suggested that major jet stretching took place in the straight electrified jet region, and further chain orientation could be gained by the subsequent process of jet whipping. By changing PHB concentrations from 5 to 15 wt%, the solution viscosity (ηo) was increased from 100 to 4900 cP, whereas the surface tension and solution conductivity remained unchanged; it provided a good model solution to exclusively reveal the ηo effect on the electrospinning process. Our results showed that the ηo-dependence of dj and df also followed simple scaling laws: dj-ηo0.06, and df-ηo0.39, with a prefactor depending on the processing variables, mainly the flow-rate. Regardless of the PHB concentrations used, the obtained PHB fibers showed a similar crystallinity fraction of ca. 0.63 and possession of major α-crystals together with a small amount of β-crystals with zigzag chain conformation.  相似文献   

12.
Heat transfer from a vertical grid jet within a 2 ft diameter and 4 ft deep fluidized bed of cracking catalyst was studied. The test nozzle diameter was varied from ¼ to 1 in. and the nozzle velocity from 50 to 250 ft /sec which is within the range of industrial practice. The axial temperature data have been related to a Froude, a Reynolds and a Nozzle number: In (δT/δTo) = –58.1 Fr?0.562 No1.08 Re?0.112 A simple jet quenching model yielded heat transfer coefficients between the fluid bed and grid jet which ranged from 300 to 1200 Btu/ft2 hr.o F.  相似文献   

13.
Jet-spouted beds characterised by high velocity gas jets (above 1.7 Umsl), and shallow bed depths H0 of around 2 D1 were investigated on laboratory scale beds and industrial scale beds and the results obtained thereof are correlated and presented in this work. Compared with the classical spouted beds, important differences in bed structure, solid movements and basic hydrodynamic characteristics were observed. The minimum spouting velocity, bed voidage and pressure drop during stabilized spouting are described in terms of dimensionless equations. Bed expansion was used as the basis for the classification of different jet-spouting regimes (incoherent spouting, fast spouting, pneumatic conveying) and changes in the slope of the bed expansion curve are correlated with regime changes. This classification could be useful in the optimization of industrial scale jet-spouted beds. A typically applicable regime of fast spouting was identified.  相似文献   

14.
In this study, physical and thermodynamic properties of poly(2-phenyl-1,3-dioxolane-4-yl-methyl-methacrylate-co-styrene) (PDMMA-ST) were investigated by using inverse gas chromatography. Two groups of solvents with different chemical natures and polarities were used to obtain information about PDMMA-ST-solvent interactions: alcohols and alkanes. The specific retention volume (Vg 0), the sorption enthalpy (ΔH1 S), sorption free energy (ΔG1 S), sorption entropy (ΔS1 S), the weight fraction activity coefficients of solute probes at infinite dilution (Ω 1 ), and Flory-Huggins interaction parameters (χ 12 ) between polymer and solvents were determined for the interactions of PDMMA-ST with alcohols and alkanes by inverse gas chromatography in the temperature range of 333–473 K. Also, the solubility parameters of PDMMA-ST at infinite dilution were found by plotting the graph of [(δ 1 2/RT) - χ 12 /V1] versus solubility parameters, δ 1, of probes.  相似文献   

15.
A high rate fibre filter was used as a pre-treatment to seawater reverse osmosis (SWRO) to reduce membrane fouling. Seawater was drawn from Chowder Bay where the Sydney Institute of Marine Science, Australia is located. A lab-scale fibre filter with a height of 1000 mm and a diameter of 30 mm was used in conjunction with in-line coagulation. The effect of operating the fibre filter with different packing densities (105, 115 kg/m3) and filtration velocities (40, 60 m/h) was investigated in terms of silt density index (SDI10), modified fouling index (MFI), pressure drop (ΔP), turbidity and molecular weight distribution (MWD). The use of in-line coagulation improved the performance of fibre filter as measured by the MFI and SDI. Regardless of filtration velocity and packing density the MFI and SDI10 values remained low as did the turbidity until the end of the filtration run. The MWD analysis showed the removal efficiencies of organic materials like biopolymers, fulvic acids, low MW acids for even experiments with the highest filtration velocity (60 m/h) and lowest packing density (105 kg/m3). This pre-treatment has a small foot print as it has the capacity of operating at a very high filtration velocity.  相似文献   

16.
Correlations for predicting characteristic velocity both above and below the critical rotor speed have been obtained under conditions with and without solute transfer. It has been found that (1) above the critical rotor speed, the characteristic velocity U0 is proportional to g/DrN2, whereas below this value a transition region exists where Uo is proportional to (g/DrN2)0.26; (2) multiple regression analysis of the experimental data of continuous phase axial mixing shows that the axial dispersion coefficient varies not only with the rotor speed and modified velocity of continuous phase but also with the velocity of dispersed phase. With varying RDC operations, the true value of Kod corrected for axial mixing changes continually between the limiting values predicted from stagnant and fully turbulent drop models. However, the highest experimental values were only 30 to 40% of those predicted by the Handlos-Baron model at the same drop Peclet number.  相似文献   

17.

A model to predict the atmospheric dry deposition velocities of particles has been developed that is similar to a model developed for the prediction of particle deposition velocities in vertical pipes. The model correlates the particle deposition velocity (Vd ) with Stokes settling velocity (V st), friction velocity (V*), dimensionless inertial deposition velocity (V + di ), and dimensionless Brownian diffusion deposition velocity (V + dd ). V + di   相似文献   

18.
The mass transport characteristics of a porous, rotating cylinder electrode (RCE, 1.0 cm diameter; 0.5, 0.9 or 1.2 cm long; 1.25, 2.25, 3.00 cm3 overall volume; 250-2000 rpm speed) fabricated from reticulated vitreous carbon (RVC, 60 ppi or 100 ppi) were investigated. The deposition of copper from an acid sulfate electrolyte (typically, deoxygenated 1 mM CuSO4 in pH 2, 0.5 M Na2SO4 at 298 K) was used as a test reaction. The effect of a jet flow of electrolyte towards the electrode and the introduction of polypropylene baffles in the electrochemical cell were studied at controlled rotation rates of the RCE. The product of mass transport coefficient and volumetric electrode area (kmAe) is related to the rotation speed of the electrode. For the 60 ppi RVC RCE, the jet electrolyte flow (3.5 cm3 s−1) enhanced the mass transport rates by a factor of 1.46 at low rotation speeds; this factor was reduced to 1.08 at high rotation speeds. For a 100 ppi electrode, the enhanced mass transport decreased from 1.26 to 1.03 at low and high rotation rates, respectively. Under the experimental conditions, baffles showed little effect on the mass transport rates to the RVC RCE. Mass transport to jet flow at an RVC RCE is compared to other RCEs using dimensionless group correlation.  相似文献   

19.
A numerical computational method to solve the problems of mass transport to the impinged surface of a wall-jet electrode reactor is put forward, thus providing the necessary tool for a quantitative electrochemical investigation of the mechanism of electrode processes, using a wall-jet electrode reactor as a hydrodynamic electrode system. The computational method is based on a second order-correct implicit finite difference approach and a coordinate transformation making a simple Cartesian space discretization compatible with efficient computing, thus allowing the computations to be performed on a personal computer. The computational approach is demonstrated through calculation of a single step chronoamperometric transient for a simple one electron transfer reaction and shown to be accurate by comparing the computed with experimentally determined current transients using as a model reaction the reduction of ferricyanide ions at a platinum electrode surface from a 0.01 m K3Fe(CN)6-0.01 m K4Fe(CN)6 solution containing l m KCl as supporting electrolyteList of symbols a nozzle diameter (m) - C i concentration of electroactive species i (mol m–3) - C i normalized concentration of electroactive species i - D i diffusion coefficient of the electroactive species i (m2 s–1) - E electrode potential (V vs SCE) - E 0 equilibrium potential (V vs SCE) - F Faraday's constant (C mol–1) - dimensionless parameter, describing the distance normal to the impinged electrode - H distance between the working electrode and the tip of the nozzle (m) - I electrode current (A) - k r constant linking the typical velocity of the wall-jet to the mean velocity in the nozzle - M flux of exterior momentum flux - v kinematic viscosity (m2 s–1) - r distance along the impinged electrode in cylindrical pole coordinates having their origin at the intersection of the jet axis and the electrode surface - R radius of the impinged electrode (m) - dimensionless time - t time (s) - v I velocity component along the impinged electrode (m s–1) - v Z velocity component normal to the impinged electrode (m s–1) - V f volume flow rate (m–3 s–1) - dimensionless parameter, describing the distance normal to the impinged electrode - z distance normal to the impinged electrode in cylindrical pole coordinates having their origin at the intersection of the jet axis and the electrode surface (m)  相似文献   

20.
A wall jet electrode reactor possessing a laminar flow regime, suitable for mechanistic studies, is reported. This reactor is different from those described in the literature in the size of its working electrode surface area. The reactor is evaluated by means of mass transport-limited current measurements using as a model reaction the reduction of ferricyanide ions at a platinum electrode surface from a 0.01 m K3Fe(CN)6-0.01 m K4Fe(CN)6 solution containing 1 m KCl as supporting electrolyte. The dependence of the mass transport-limited current on the crucial reactor parameters — the volume flow rate V f (m3 s–1), the nozzle diameter a (m) and the radius of the working electrode R (m) — is established and verified by theoretical predictions. The reactor is shown to have the desired wall jet hydrodynamics for: 1.6 × 10–6 V f 4.3 × 10–6 m3 s–1, 1.5 × 10–3 a 3 × 10–3 m and 1.5 × 10–2 R 2 × 10–2 m.List of symbols a nozzle diameter (m) - C A concentration of A in the bulk (mol m–3) - D A diffusion coefficient of A (m2 s–1) - F Faraday's constant (C mol–1) - dynamic viscosity (gm–1 s–1) - H distance between the working electrode and the tip of the nozzle (m) - I lim mass-transport-limited current (A) - k r constant linking the typical velocity of the wall-jet to the mean velocity in the nozzle - v kinematic viscosity (m2 s–1) - n number of electrons exchanged - density (g m–3) - R radius of the working electrode (m) - t time (s) - V f volume flow rate (m–3 s–1)  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