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 296 毫秒
1.

The dairy processing industry in India, on an average basis, involves an extensive amount of thermal and electrical energy consumption, i.e. 2.51 × 105 kW MT−1 and 1.44 × 105 kW MT−1, respectively, for an installed milk food processing capacity of 1.21 × 105 TPD. However, energy consumption spectrum depends upon the level of automation along with better utilisation of utility resources. The global ultra-high-temperature (UHT) pasteurised milk trade was valued at € 52.29 billion in 2012 and is expected to reach € 114.38 billion by 2019–2020. In the present work energy, exergy and exergoeconomic evaluation of ultra-high-temperature milk pasteurisation plant have been considered. The overall energy efficiency and efficiency pertaining to executable potential of energy in UHT Milk Processing Unit were reported to be 86.36% and 53.02%. The specific exergy destruction and specific exergy improvement potential were estimated to be 219.23 kJ kg−1 and 137.60 kJ kg−1, respectively. The highest possible retrievable exergy potential of the plant was associated with heating coil, i.e. 158.98 kW, followed by homogeniser (54.62 kW), which pinpointed towards the possibility of huge technical improvement. The processing cost was enumerated to be highest for heating coil (rk: 38.35%) followed by regeneration-1 (rk: 23.40%). Further, the total operating cost rate associated with thermodynamic deficiencies of subunits was estimated to be highest for heating coil (4859.82 € H−1) followed by regenerator-2 (1264.88 € H−1) and homogeniser (1187.14 € H−1). The broad survey of thermoeconomic indices of subunits indicated that the level of exergetic destruction was far more on higher side.

  相似文献   

2.

In this work, two newly sensitive and selective Al(III)-modified carbon paste electrodes (MCPEs) were developed based on diphenylcarbazone (DPC) modifier mixed with tricresyl phosphate plasticizer and either graphite powder (electrode I) or graphite powder mixed with graphene (electrode II). The potentiometric performance characteristics of the two electrodes were scrutinized and discussed. The proposed sensors showed a high electrochemical response in the linear concentration range of 1.0 × 10−6 to 1.0 × 10−2 mol L−1 with a good Nernstian slopes of 20.12 ± 0.30 mV decade−1 and 20.63 ± 0.66 mV decade−1 and limits of detection of 9.0 × 10−7 and 8.5 × 10−7 mol L−1 for electrode (I) and electrode (II), respectively. Both electrodes showed a fast response time and reasonable thermal stability. The potentiometric response of the DPC-based electrodes was independent on the pH of the tested solutions in ranges of 2.5–5 and 2.5–5.5 for electrode (I) and electrode (II), respectively. The two electrodes can be also used in partially non-aqueous medium containing up to 20% (v/v) acetone or methanol with no significant changes in the working concentration ranges or the slopes. The proposed electrodes showed fairly good discriminating ability toward Al(III) ions in comparison with many other metal ions. The electrodes were applied successfully for Al(III) ions determination in drainage water, spiked tap water and pharmaceutical preparation samples. Furthermore, the electrode surfaces were characterized using energy-dispersive X-ray (EDX) and scanning electron microscopic (SEM) as surface characterization techniques and Fourier Transform Infrared (FT-IR) technique to confirm the interaction between Al(III) and DPC.

  相似文献   

3.

Isothermal and dynamic differential scanning calorimetry (DSC) was exploited to study the curing behavior of diglycidyl ether bisphenol-A epoxy resin with various combining ratios of dicyandiamide (DICY) and nadic methyl anhydride (NMA). Curves of prepared samples indicated that the enthalpy of the reaction decreased with increasing the molar ratios (NMA/DICY) up to 40% after which an exothermic peak peculiar to the effect of anhydride appeared at a higher temperature. The curing behavior examination of the samples containing the aforementioned molar ratio of NMA/DICY (= 40%) was carried out using isothermal condition at different temperatures (130–145 °C) and dynamic condition DSC at various heating rates (2.5–20 °C min−1). Under the isothermal condition, by constructing a master curve, the values of activation energy (Ea) and pre-exponential factor (A) were calculated 89.3 kJ mol−1 and 1.2 × 10+9 s−1, respectively. The activation energy of the curing reactions in a dynamic mode was obtained 85.32 kJ mol−1 and 88.02 kJ mol−1 using Kissinger and Ozawa methods, respectively. Likewise, pre-exponential factors were also calculated 3.35 × 10+8 and 7.4 × 10 +8 s−1, respectively. The overall order of reaction for both conditions was found to be a value around 3.

  相似文献   

4.
The reactions between OH radicals and hydrogen halides (HCl, HBr, HI) have been studied between 298 and 460 K by using a discharge flow-electron paramagnetic resonance technique. The rate constants were found to be kHCl(298 K) = (7.9 ± 1.3) × 10−13 cm3 molecule−1 s−1 with a weak positive temperature dependence, kHBr (298-460 K) = (1.04 ± 0.2) × 10−11 cm3 molecule−1 s−1, and kHI(298 K) = (3.0 ± 0.3) × 10−11 cm3 molecule−1 s−1, respectively. The homogeneous nature of these reactions has been experimentally tested.  相似文献   

5.
The overall rate constants for H-abstraction (kH) from tetrahydrofuran and D-abstraction (kD) from fully deuterated tetrahydrofuran by chlorine atoms in the temperature range of 298-547 K were determined. In both cases, very weak negative temperature dependences of the overall rate constants were observed, described by the expressions: kH = (1.55 ± 0.13) × 10−10 exp(52 ± 28/T) cm3 molecule−1 s−1 and kD = (1.27 ± 0.25) × 10−10exp(55 ± 62/T) cm3 molecule−1 s−1. The experimental results show that the value of the kinetic isotope effect (kH/kD), amounting to 1.21 ± 0.10, is temperature independent at 298-547 K.  相似文献   

6.
Kinetics of the reaction of Cl atoms with methanol has been investigated at 2 Torr total pressure of helium and over a wide temperature range 225-950 K, using a discharge flow reactor combined with an electron impact ionization quadrupole mass spectrometer. The rate constant of the reaction Cl + CH3OH → products (1) was determined using both absolute measurements under pseudo-first order conditions, monitoring the kinetics of Cl-atom consumption in excess of methanol and relative rate method, k1 = (5.1 ± 0.8) × 10−11 cm3 molecule−1 s−1, and was found to be temperature independent over the range T = 225-950 K. The rate constant of the reaction Cl + Br2 → BrCl + Br (3) was measured in an absolute way monitoring Cl-atom decays in excess of Br2: k3 = 1.64 × 10−10 exp(34/T) cm3 molecule−1 s−1 at T = 225-960 K (with conservative 15% uncertainty). The experimental data for k3 can also be adequately represented by the temperature independent value of k3 = (1.8 ± 0.3) × 10−10 cm3 molecule−1 s−1. The kinetic data from the present study are compared with previous measurements.  相似文献   

7.
In this study, the kinetics and mechanism of UV/O3 synergistic oxidative digestion of dissolved organic phosphorus (DOP) were investigated, focusing on the ozone direct oxidation and hydroxyl radical oxidation parts of glufosinate and triphenyl phosphate (TPhP). The p-chlorobenzoic acid (p-CBA) was selected as the probe compound, and two kinds of reaction kinetic models were proposed by competitive kinetic method with Rct according to the different scale of rate constants of hydroxyl radical oxidation. Under the condition of weakly alkaline (pH = 9.0) and weakly acidic environment (pH = 5.0), the second-order rate constants of glufosinate and TPhP was determined indirectly to be ko3/glufosinate = (2.903 ± 0.247)M−1s−1 and ko3/TPhP = (3.307 ± 0.204) M−1s−1 by ozone direct oxidation, and k·OH/glufosinate = (1.257 ± 1.031) × 109 M−1s−1 and k·OH/TPhP = (7.120 × 108 ± 0.963) M−1s−1 by hydroxyl radical oxidation, respectively. The comparison of the contribution levels of the two parts to the digestion process showed that the contribution levels in the digestion of glufosinate and TPhP processes both the contribution of ·OH were higher than those of ozone, 86.3% and 72.6%, respectively.  相似文献   

8.
Fourier transform infrared (FTIR) smog chamber techniques were used to investigate the atmospheric chemistry of the isotopologues of methane. Relative rate measurements were performed to determine the kinetics of the reaction of the isotopologues of methane with OH radicals in cm3 molecule−1 s−1 units: k(CH3D + OH) = (5.19 ± 0.90) × 10−15, k(CH2D2 + OH) = (4.11 ± 0.74) × 10−15, k(CHD3 + OH) = (2.14 ± 0.43) × 10−15, and k(CD4 + OH) = (1.17 ± 0.19) × 10−15 in 700 Torr of air diluent at 296 ± 2 K. Using the determined OH rate coefficients, the atmospheric lifetimes for CH4–xDx (x = 1–4) were estimated to be 6.1, 7.7, 14.8, and 27.0 years, respectively. The results are discussed in relation to previous measurements of these rate coefficients.  相似文献   

9.

In this research, electrospun polycaprolactam nanofibers were collected on a fine stainless steel mesh sheet without a binder, and a layer of conductive polyaniline was chemically deposited on the nanofibers. The polyaniline immobilized on the polycaprolactam nanofibers provided high electrical conductivity, acceptable mechanical stability, and a large surface area. This assembly was then used as a working electrode in electrochemically controlled solid-phase microextraction (EC-SPME), a fast and environmentally friendly method. The polymer layers were characterized by SEM and FTIR techniques. Significant factors affecting the EC-SPME efficiency were investigated, including the desorption conditions, the sorbent used, the pH of the sample solution, the extraction voltage, the extraction time, and the ionic strength. Under the optimum conditions, the limits of detection and quantification for the target analytes were 0.9–1.8 μg L−1 and 3.0–6.1 μg L−1, respectively. The linear dynamic range was 5–2000 μg L−1, with R2 > 0.993. The method was coupled with HPLC analysis and applied to the determination of angiotensin ΙΙ receptor antagonists (ARA-ΙΙs) in human plasma, and relative recoveries of 91.1–104.3% with RSDs of ≤8.3% were obtained.

  相似文献   

10.

Four types of undisturbed soils around the Es-Salam reactor (Algeria) were used to evaluate the sorption behavior of strontium. The batch study was carried out under different experimental conditions. The kinetics were well fited by pseudosecond order model. Soils’s activation energies were 12.37, 14.76, 15.5 and 16.17 kJ mol−1, corresponding to ion-exchange-type sorption. Sorption was exothermic (ΔH° < 0), spontaneous (ΔG° < 0) and favorable at low temperature. Competing cations, particularly Ca2+ reduce the Sr adsorption. Desorption reaction showed a higher value of Sr in the easily extractible phase indicating a relative availability of the element.

  相似文献   

11.
The spin-forbidden dissociation reaction of the N2O(X1Σ+) ground state has been investigated by both quantum calculations and experiments. Ab initio prediction at the CCSD(T)/CBS(TQ5)//CCSD(T)/aug-cc-pVTZ+d level of theory gave the crossing point (MSX) energy at 60.1 kcal/mol for the N2O(X1Σ+) → N2() + O(3P) transition, in good agreement with published data. The T- and P-dependent rate coefficients, k1(T,P), for the nonadiabatic thermal dissociation predicted by nonadiabatic Rice-Ramsperger-Kassel-Marcus (RRKM) calculations agree very well with literature data. The rate constants at the high- and low-pressure limits, k1 = 1011.90 exp (−61.54 kcal mol−1/RT) s−1 and k1o = 1014.97 exp(−60.05 kcal mol−1/RT) cm3 mol−1 s−1, for example, agree closely with the extrapolated results of Röhrig et al. at both pressure limits. The second-order rate constant (k1o) is also in excellent agreement with our result measured by FTIR spectrometry in the present study for the temperature range of 860-1023 K as well as with many existing high-temperature data obtained primarily by shock-wave heating up to 3340 K. Kinetic modeling of the NO product yields measured at long reaction times in the present work also allowed us to reliably estimate the rate constant for reaction (3), O + N2O → N2 + O2, based on its strong competition with the NO formation from reaction (2) which has been better established. The modeled values of k3 confirmed the previous finding by Davidson et al. with significantly smaller values of A-factor and activation energy than the accepted ones. A least-squares analysis of both sets of data gave k3 = 1012.22 ± 0.04 exp[− (11.65 ± 0.24 kcal mol−1/RT)] cm3 mol−1 s−1, covering the wide temperature range of 988-3340 K.  相似文献   

12.
Rate coefficients, k1, for the gas-phase OH radical reaction with the heterocyclic ether C4H4O (1,4-epoxybuta-1,3-diene, furan) were measured over the temperature range 273–353 K at 760 Torr (syn. air). Experiments were performed using: (i) the photochemical smog chamber THALAMOS (thermally regulated atmospheric simulation chamber, IMT NE, Douai-France) equipped with Fourier Transform Infrared (FTIR) and Selected Ion Flow Tube Mass Spectrometry (SIFT-MS) detection methods and (ii) a photochemical reactor coupled with FTIR spectroscopy (PCR, University of Crete, Greece). k1(273–353 K) was measured using a relative rate (RR) method, in which the loss of furan was measured relative to the loss of reference compounds with well-established OH reaction rate coefficients. k1(273–353 K) was found to be well represented by the Arrhenius expression (1.30 ± 0.12) × 10−11 exp[(336 ± 20)/T] cm3 molecule−1 s−1, with k1(296 K) measured to be (4.07 ± 0.32) × 10−11 cm3 molecule−1 s−1. The k1(296 K) and pre-exponential quoted error limits are 2σ and include estimated systematic errors in the reference rate coefficients. The observed negative temperature dependence is consistent with a reaction mechanism involving the OH radical association to a furan double bond. Quantum mechanical molecular calculations show that OH addition to the α-carbon (ΔHr(296 K) = −121.5 kJ mol−1) is thermochemically favored over the β-carbon (ΔHr(296 K) = −52.9 kJ mol−1) addition. The OH-furan adduct was found to be stable over the temperature range of the present measurements. Maleic anhydride (C4H2O3) was identified as a minor reaction product, 3% lower-limit yield, demonstrating a non-ring-opening active reaction channel. The present results are critically compared with results from previous studies of the OH + furan reaction rate coefficient. The infrared spectrum of furan was measured as part of this study and its estimated climate metrics are reported.  相似文献   

13.
The rate constants of the reactions of DO2 + HO2 (R1) and DO2 + DO2 (R2) have been determined by the simultaneous, selective, and quantitative measurement of HO2 and DO2 by continuous wave cavity ring-down spectroscopy (cw-CRDS) in the near infrared, coupled to a radical generation by laser photolysis. HO2 was generated by photolyzing Cl2 in the presence of CH3OH and O2. Low concentrations of DO2 were generated simultaneously by adding low concentrations of D2O to the reaction mixture, leading through isotopic exchange on tubing and reactor walls to formation of low concentrations of CH3OD and thus formation of DO2. Excess DO2 was generated by photolyzing Cl2 in the presence of CD3OD and O2, small concentrations of HO2 were always generated simultaneously by isotopic exchange between CD3OD and residual H2O. The rate constant k1 at 295 K was found to be pressure independent in the range 25–200 Torr helium, but increased with increasing D2O concentration k1 = (1.67 ± 0.03) × 10−12 × (1 + (8.2 ± 1.6) × 10−18 cm× [D2O] cm−3) cm3 s−1. The rate constant for the DO2 self-reaction k2 has been measured under excess DO2 concentration, and the DO2 concentration has been determined by fitting the HO2 decays, now governed by their reaction with DO2, to the rate constant k1. A rate constant with insignificant pressure dependence was found: k2 = (4.1 ± 0.6) × 10−13 (1 + (2 ± 2) × 10−20 cm× [He] cm−3) cm3 s−1 as well as an increase of k2 with increasing D2O concentration was observed: k2 = (4.14 ± 0.02) × 10−13 × (1 + (6.5 ± 1.3) × 10−18 cm3 × [D2O] cm−3) cm3 s−1. The result for k2 is in excellent agreement with literature values, whereas this is the first determination of k1.  相似文献   

14.
The kinetics of the reactions of Br2 and NO2 with ground state oxygen atoms have been studied over a wide temperature range, T = 220-950 K, using a low-pressure flow tube reactor coupled with a quadrupole mass spectrometer: O + NO2 → NO + O2 (1) and O + Br2 → Br + BrO (2). The rate constant of reaction (1) was determined under pseudo–first-order conditions, either monitoring the kinetics of O-atom or NO2 consumption in excess of NO2 or of the oxygen atoms, respectively: k1 = (6.1 ± 0.4) × 10−12 exp((155 ± 18)/T) cm3 molecule−1 s−1 (where the uncertainties represent precision at the 2σ level, the estimated total uncertainty on k1 being 15% at all temperatures). The temperature dependence of k1, found to be in excellent agreement with multiple previous low-temperature data, was extended to 950 K. The rate constant of reaction (2) determined under pseudo–first-order conditions, monitoring the kinetics of Br2 consumption in excess of O-atoms, showed upward curvature at low and high temperatures of the study and was fitted with the following three-parameter expression: k2 = 9.85 × 10−16 T1.41 exp(543/T) cm3 molecule−1 s−1 at T = (220-950) K, which is recommended from the present study with an independent of temperature conservative uncertainty of 15% on k2.  相似文献   

15.
The kinetics of the reaction of H-atom with carbonyl sulfide (OCS) has been investigated at nearly 2 Torr total pressure of helium over a wide temperature range, T = 255–960 K, using a low-pressure discharge flow reactor combined with an electron impact ionization quadrupole mass spectrometer. The rate constant of the reaction H + OCS → SH + CO (1) was determined under pseudo-first order conditions, monitoring the kinetics of H-atom consumption in excess of OCS, k1 = 6.6 × 10−13 × (T/298)3 × exp(−1150/T) cm3 molecule−1 s−1 (with estimated total uncertainty on k1 of 15% at all temperatures). Current measurements of k1 at intermediate temperatures (520–960 K) appear to reconcile previous high and low temperature data and allow the above expression for k1 to be recommended for use in the extended temperature range between 255 and 1830 K with a conservative uncertainty of 20%.  相似文献   

16.

ColiSense, an early warning system developed for Escherichia coli detection, is assessed using environmental samples. The system relies on the detection of β-glucuronidase (GUS), a biomarker enzyme for E. coli. In contrast with other rapid GUS-based methods, ColiSense is the only method that uses 6-chloro-4-methyl-umbelliferyl-β-d-glucuronide (6-CMUG) as a fluorogenic substrate. The system measures a direct kinetic response of extracted GUS, and the detection was carried out in the absence of particles or bacteria. It is necessary to evaluate the system with environmental samples to establish the relationship between faecal indicator bacteria E. coli and the response measured by the ColiSense. This paper presents the results of tests carried out with the ColiSense system for 2 trials, one conducted with freshwater samples collected from rivers in the Dublin area and a second conducted with seawater samples from coastal areas collected over the bathing season. A positive linear correlation was found between E. coli (MPN 100 mL−1) and ColiSense response (R2 = 0.85, N = 125, p < 0.01) for the seawater sample. A ColiSense response threshold was identified as 0–1.8 pmol min−1 100 mL−1, equivalent to 0–500 E. coli 100 mL−1. Using this threshold, 96.8% of the samples were correctly classified as being above or below 500 E. coli 100 mL−1 by the ColiSense system. Results presented demonstrate that the ColiSense system can be used as an early warning tool with potential for active management of bathing areas by providing results in 75 min from sample collection.

  相似文献   

17.
The rate coefficients of the reactions of OH radicals and Cl atoms with three alkylcyclohexanes compounds, methylcyclohexane (MCH), trans‐1,4‐dimethylcyclohexane (DCH), and ethylcyclohexane (ECH) have been investigated at (293 ± 1) K and 1000 mbar of air using relative rate methods. A majority of the experiments were performed in the Highly Instrumented Reactor for Atmospheric Chemistry (HIRAC), a stainless steel chamber using in situ FTIR analysis and online gas chromatography with flame ionization detection (GC‐FID) detection to monitor the decay of the alkylcyclohexanes and the reference compounds. The studies were undertaken to provide kinetic data for calibrations of radical detection techniques in HIRAC. The following rate coefficients (in cm3 molecule−1 s−1) were obtained for Cl reactions: k(Cl+MCH) = (3.51 ± 0.37) × 10–10, k(Cl+DCH) = (3.63 ± 0.38) × 10−10, k(Cl+ECH) = (3.88 ± 0.41) × 10−10, and for the reactions with OH radicals: k(OH+MCH) = (9.5 ± 1.3) × 10–12, k(OH+DCH) = (12.1 ± 2.2) × 10−12, k(OH+ECH) = (11.8 ± 2.0) × 10−12. Errors are a combination of statistical errors in the relative rate ratio (2σ) and the error in the reference rate coefficient. Checks for possible systematic errors were made by the use of two reference compounds, two different measurement techniques, and also three different sources of OH were employed in this study: photolysis of CH3ONO with black lamps, photolysis of H2O2 at 254 nm, and nonphotolytic trans‐2‐butene ozonolysis. For DCH, some direct laser flash photolysis studies were also undertaken, producing results in good agreement with the relative rate measurements. Additionally, temperature‐dependent rate coefficient investigations were performed for the reaction of methylcyclohexane with the OH radical over the range 273‐343 K using the relative rate method; the resulting recommended Arrhenius expression is k(OH + MCH) = (1.85 ± 0.27) × 10–11 exp((–1.62 ± 0.16) kJ mol−1/RT) cm3 molecule−1 s−1. The kinetic data are discussed in terms of OH and Cl reactivity trends, and comparisons are made with the existing literature values and with rate coefficients from structure‐activity relationship methods. This is the first study on the rate coefficient determination of the reaction of ECH with OH radicals and chlorine atoms, respectively.  相似文献   

18.
19.
Pure and doped samples of potassium bromate (KBrO3) were subjected to precompression and their thermal decomposition kinetics was studied by thermogravimetry at 668 K. The samples decomposed in two stages governed by the same rate law (contracting square equation), but with different rate constants, k 1 (for a α ≤ 0.45) and k 2 (for α ≥ 0.45), as in the case of uncompressed samples. The rate constants k 1 and k 2 decreased dramatically on precompression, the decrease being higher for doped samples. Cation dopants (Ba2+, Al3+) caused more desensitization effect than the anion dopants ( \textSO4 2- {\text{SO}}_{4}{}^{ 2- } , PO4 3−) of the same magnitude of charge and concentration. The results favor ionic diffusion mechanism proposed earlier on the basis of doping studies.  相似文献   

20.
Qiao  Dan  Wang  Yue  Li  Fan  Wang  Daya  Yan  Baijun 《Journal of Thermal Analysis and Calorimetry》2019,137(2):389-397

Controlling the conditions of the oxygen partial pressure and temperature to prepare the WO2.72 (W18O49) via reduction was possible through thermodynamic consideration. WO2.72 was synthesized via heating to 1073 K in 5% H2–95% Ar mixture gas flow from ammonium tungstate which was prepared by hydrothermal process. With the reducing prolonging time, the products were changed from WO2.72 to WO2 and then metal W. Thermogravimetric (TG) analysis showed ammonium tungstate decomposed completely to WO3 at 773 K. Isothermal reductions using TG analysis were carried out at 905 K, 925 K, 945 K and 973 K in 5% H2–95% Ar mixture gas flow, respectively. The whole reduction from WO3 to WO2.72 divided into three parts: initial nucleation and growth stage, final interfacial reaction stage and intermediate stage, was controlled jointly by both mechanisms. Fitting results showed that the initial stage obey the one-dimensional Avrami–Erofeev equation, the apparent activation energy was 132.7 ± 1.1 kJ mol−1 and the pre-exponent factor was 4.82 × 105 min−1; the final stage expressed by 2-dimensional interfacial reaction, the apparent activation energy was 144.0 ± 2.1 kJ mol−1 and the pre-exponent factor was 3.20 × 105 min−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