You are on page 1of 6

Preprints of the 8th IFAC Symposium on Advanced Control of Chemical Processes The International Federation of Automatic Control Furama

Riverfront, Singapore, July 10-13, 2012

Modelling and Simulation of Counter-Current and Confined Jet Reactors for Continuous Hydrothermal Flow Synthesis of Nano-materials
Cai Y. Ma*, Xue Z. Wang*, Chris J. Tighe** and Jawwad A. Darr** *Institute of Particle Science and Engineering, School of Process, Environmental and Materials Engineering, University of Leeds, Leeds LS2 9JT, United Kingdom (x.z.wang@leeds.ac.uk) **Department of Chemistry, University College London, Christopher Ingold Laboratories, 20 Gordon Street, London WC1H 0AJ, United Kingdom Abstract: Computational fluid dynamics is applied to a comparative study of a counter-current reactor and a confined jet reactor for continuous hydrothermal flow synthesis of nanomaterials under supercritical water conditions. The fluid flow and heat transfer variables including velocity and temperature profiles in both reactor configurations are simulated using ANSYS Fluent package. The tracer concentration profiles are also modelled via solving species equations from which the mixing behaviour in the reactors is investigated. The predicted temperatures are found to be in good agreement with experimental data. The simulation also provides suggestions to improving the reactor designs and process control. Key words: Reactor modeling, reactor simulation, reactor control, nanomaterials, hydrothermal synthesis 2. MATHEMATICAL FORMULATION 2.1 Hydrodynamic and mixing models The three-dimensional continuity, momentum and enthalpy equations based on time-averaged quantities obtained from Reynolds averaging of the instantaneous equations are numerically solved to obtain hydrodynamic and heat transfer profiles. The most commonly used k - two-equation model was used in this study with the turbulent kinetic energy k and its dissipation rate being calculated with standard empirical constants (Launder et al., 1975). Simulation of mixing was carried out by introducing a secondary liquid as an inert tracer into the primary liquid in the reactors. The temporal and spatial distributions of the tracer concentration were obtained from the solution of the Reynolds-averaged species transport equation (Eq. 1): (1) where is the mixture density, the effective diffusion coefficient of species i is , ui is the mixture velocity, Yi represents the species mass fraction, i the species diffusion coefficient, t the turbulent eddy viscosity, and Sct the turbulent Schmidt number. 2.2 Thermodynamic properties Considering the need of both efficiency and accuracy, the water properties obtained from National Institute of Standard and Technology (2009) using the 1995 IAPWS (International Association for the Properties of Water and Steam) formulation (Wagner and Pru, 2002) were piece-wise curvefitted in polynomial forms over several temperature ranges at a pressure of 24.1 MPa (Ma et al., 2011), and used to calculate thermodynamic properties. For diffusion coefficients in the mixing study, the correlation from Liu and Macedo (1995) was used to estimate the water diffusivity.

1. INTRODUCTION Continuous hydrothermal flow synthesis (CHFS) is an attractive process for manufacture of metal oxide nanomaterials. The process is relatively green since it uses water rather than organic solvents; and it is easy to control, avoiding batch to batch variation in product quality (Adschiri et al., 1992, Boldrin et al., 2006, Tighe et al., 2011). Optimisation of the mixing condition and temperature and velocity profiles are critical to process operation and control, and product quality. Computational fluid dynamics (CFD) modelling offers a useful tool to the evaluation and optimisation of designs, and to the scale up and control of the process. Liu and Fox (2006) studied continuous impinging jet reactors using CFD under atmospheric conditions. Antisolvent crystallisation processes were investigated with the similar reactor configuration by combining CFD and population balance (PB) techniques (Woo et al. 2009). There has also been previous research on the application of CFD to hydrothermal synthesis process (Aimable et al., 2009, Demoisson et al., 2011, Kawasaki et al., 2010, Lester et al., 2006, Moussiere et al., 2007, Narayanana et al., 2008, SierraPallares et al., 2011), but previous work is limited because the majority did not validate simulation results against measurements or validated insufficiently. In addition, some simulations were conducted at below supercritical conditions. In this paper, CFD models for a counter-current reactor and a confined jet mixer of a CHFS system were developed. Experiments were performed under various operating conditions for the reactors to collect data for validation of simulation results. The study has focused on the investigation of the effect of operating conditions and reactor configurations on the process behaviour including flow field, temperature and concentration. Such knowledge is needed for optimising reactor design, successful scaling-up and also process control.
IFAC, 2012. All rights reserved. 874

8th IFAC Symposium on Advanced Control of Chemical Processes Furama Riverfront, Singapore, July 10-13, 2012

3.

EXPERIMENTS

Figure 1 shows the schematic of the CHFS process. Deionised water was pumped through a heated, helical coil using pump P1 with a fixed pressure of 24.1 MPa and a constant flowrate in the range 10 to 25 mL/min. Since water density, hence volumetric flowrate, varies with temperature, for comparison purposes, in this study all volumetric inlet flowrates quoted are based on the density of water at atmospheric pressure and at 15oC. A cartridge heater and a band heater were used to heat the deionised water to a required temperature of between 350 and 450 oC. An aqueous precursor and an aqueous base were pumped using Pumps P2 and P3, respectively, at 24.1 MPa with a fixed flowrate in the range 5 to 10 ml/min from each pump. The fluids from P2 and P3 were then mixed before entering the reactor. For counter current reactor, streams from P2 and P3 mix first before entering the reactor to meet supercritical water. For the confined jet mixer, the fluids from P2 and P3 were mixed thoroughly and the resulted mixture was then split equally into two streams before entering the mixer through two inlets. The streams from the two inlets mix with the stream from P1 and react, generating nano-particles. The slurry containing nano-particles then left the reactor to enter a heat exchanger for rapid cooling before being collected as the product. In order to concentrate the study on fluid flow and heat transfer, the mixture stream from P2 and P3 (Fig. 1) was mimicked by deionised water during temperature profiling experiments and CFD modelling, though studies are being carried out to combine CFD with PB modelling for the prediction of flow field, heat transfer phenomenon and particle size distribution and also to integrate with a control procedure for producing nanoparticles with desired particle size/shape distributions. Table 1 lists the operating conditions.

Table 1. Operating conditions for experiments and simulations (inlet temperature of precursor stream = 20oC, and operating pressures of supercritical water and precursor streams = 24.1 MPa)
Run number CCR-1 CCR-2 CCR-3 CCR-4 CCR-5 CCR-6 CCR-7 CJM-1 CJM-2 CJM-3 CJM-4 CJM-5 CJM-6 CJM-7 Supercritical Supercritical water Precursor flowrate water Tin (oC) flowrate (mL/min) (mL/min) Counter-current reactor (CCR) 350 350 400 400 400 450 450 350 350 400 400 400 450 450 20 25 10 20 25 20 25 Confined jet mixer ( CJM) 20 25 10 20 25 20 25 20 10 10 20 10 20 10 10, 10 5, 5 5, 5 10, 10 5, 5 10, 10 5, 5

The confined jet mixer ( shown in Fig. 2(right)) consists of an inner tube (0.99 mm) inserting into a larger outer tube (4.57 mm). The superheated water flows in the inner tube to the mixing point to mix with the metal salt solution from the two precursor streams. The product leaves the mixer and enters a tubular heat exchanger. Ten long, fine J-type thermocouples were inserted into the mixer at different locations along the z direction. In both reactors, the tips of the thermocouples were floated in the bulk flow due to the purpose of measuring bulk temperatures and also difficulty to fix them onto a wall. The estimated tip position variations are within 2 mm across the tube cross-section (x-y plane).

Fig. 1 Schematic of a continuous hydrothermal flow synthesis system. P1-pump for deionised water, P2-pump for aqueous precursor, P3-pump for aqueous base. R-Reactor, BPR-Backpressue regulator

The detailed schematic diagram of a counter-current reactor is illustrated in Fig. 2(left). It consists of an inner tube (1.76 mm) inserting into a larger outer tube (7.05 mm). The superheated water flows in the inner tube to the mixing point to mix with the metal salt solution. The product stream then leaves the reactor and enters a tubular heat exchanger. Seven long, fine J-type thermocouples were inserted into the reactor at different locations along the z direction.

Fig. 2 Counter current reactor (left) and confined jet mixer (right)

4. SIMULATION 4.1 Computational details The three-dimensional domain was discretised using GAMBIT software with 6.26 105 cells for the counter5 current reactor and 2.96 10 cells for the confined jet mixer. The operating conditions used for the computational studies were taken directly from the experiments and listed in Table 1. For the study of mixing, a tracer with the same operating temperature, pressure and properties as the supercritical water
875

8th IFAC Symposium on Advanced Control of Chemical Processes Furama Riverfront, Singapore, July 10-13, 2012

was introduced into the supercritical water stream. The flowrates of the tracer and supercritical water streams were 1% and 99%, respectively, of the total flowrate used.

5. RESULTS AND DISCUSSIONS 5.1 Fluid flow and heat transfer study Figure 3 shows the velocity vectors and temperature contours around the supercritical exit region of the counter-current reactor (CCR-4 in Table 1) and confined jet mixer (CJM-4 in Table 1). For the counter-current reactor, the supercritical water stream penetrated into the up-coming precursor stream to form a recirculation zone which enhanced the mixing between the supercritical water and precursor streams. The mixture then flew through the annual section to the product exit. For the confined jet mixer, the supercritical water stream flows from the inner pipe with the precursor stream flowing in the same direction but in an annulus surrounding the inner pipe. At the exit of the supercritical water stream, the two streams form a confined co-current flow configuration, in which the supercritical water jet entrains the precursor stream and forms recirculation zone surrounding the jet (Fig. 3(top)). The penetration distances for the counter-current reactor and the confined jet mixer are about 6 and 13 times of their jet exit diameters, respectively. Simulation results (not shown here) also show that from the supercritical water inlet to the exit, the supercritical water stream has been cooled down to subcritical conditions for the counter-current reactor, which results in a much lower axial velocity (about half of the inlet axial velocity), hence longer distance for the decay of tracer mass. This leads to slow mixing between the supercritical water and the precursor streams, hence affecting nanoparticle product quality. However, for the confined jet mixer, the supercritical water stream is still in supercritical conditions at the exit point, hence producing a shorter mixing distance due to the enhanced mixing. The mixing distances are estimated as 10 and 7 times of the jet exit diameters, respectively, for the counter-current reactor and confined jet mixer. The predicted and measured temperatures along the z direction were compared in Fig. 4. As the locations of the thermocouple tips can vary within 2 mm on the x-y plane in the annual section, i.e. y = 1.5875 3.5 mm for the countercurrent reactor and y = 1.0 2.0 mm for the confined jet mixer, the predicted results at y = 1.7 and 2.9 mm for the counter-current reactor, and y = 1.0 and 1.2 mm for the confined jet mixer were plotted in Fig. 4 (top) and (bottom), respectively. 5.2 Study on the mixing behaviour Contours of tracer mass fraction at 15s after starting the transient simulations are plotted in Fig. 5. The circle points are the monitoring locations along the central line of the reactors, where the tracer concentrations were recorded during the transient simulations. The monitored tracer mass fraction distributions were normalized by their fully mixed mass fraction, c . The mixing times, defined as the time required for the local tracer mass fraction to reach 99% of c , for the four monitoring points of A1 A4 (z = 103, 104, 106 and 109 mm) were estimated from the predicted tracer mass fractions with values of 2, 2.4, 3.5 and 9s, respectively, for the counter-current reactor. For the confined jet mixer, the mixing times obtained for the monitoring points B1 B4 (z = 16.885, 17.885, 19.885 and 23 mm) are 1, 2, 2.5 and 3.5s, respectively. It is demonstrated that the mixing times for the
876

(a1)

(a2)

(b1) (b2) Fig. 3 Velocity vectors (a1, a2) and temperature (b1, b2) distributions in the supercritical water exit regions of a countercurrent reactor - CCR-4 (a1, b1) and a confined jet reactor - CJM-4 (a2, b2).

4.2 Solution method The mass, momentum and energy conservation equations and species transport equation, together with the equations for k and are solved using ANSYS Fluent software (2010) . The species equations were solved with the steady-state flow and temperature profiles obtained as the initial values. Standard SIMPLE pressure-velocity coupling was used with a second order upwind scheme being employed for the discretisation of the convection terms in the governing equations. Due to the insulation of the system, the heat loss through the outer wall of the reactor was assumed to be negligible, i.e. adiabatic boundary condition having been used. The mass inlet flow mode was used to calculate the inlet velocities of both supercritical water and precursor streams. Constant inlet temperatures for the inlet fluids were specified. A turbulent intensity of 10% and the corresponding hydrodynamic diameters were used for the inlet conditions for turbulence. The outlet flow mode was used for their corresponding exit boundaries, which specify fully developed outlet flow conditions. Standard non-slip wall boundary conditions were applied in the studies with the standard turbulent wall function. Independence tests for mesh size and convergence tolerance were carried out to eliminate their effect.

8th IFAC Symposium on Advanced Control of Chemical Processes Furama Riverfront, Singapore, July 10-13, 2012

confined jet mixer are shorter, indicating a better mixing process.

good agreements with measured data at three inlet temperatures of supercritical water stream.

A1 A2 A3 A4

B1 B2 B3

B4

Fig. 5 Contours of tracer mass fraction after 15s (left) countercurrent reactor CCR-4 with four monitoring points, A1, A2, A3 and A4 (z = 103, 104, 106 and 109 mm); (right) confined jet mixer CJM-4 with four monitoring points, B1, B2, B3 and B4 (z = 16.885, 17.885, 19.885 and 23 mm).

Fig. 4 Predicted and measured (squares) temperatures along the z direction for (top) counter-current reactor CCR-4 (dotted line y = 1.7 mm; solid line y = 2.9 mm); and (bottom) confined jet mixer CJM-4 (solid line y = 1.2 mm; dashed line y = 1.0 mm).

5.3 Effect of supercritical water inlet temperatures under balanced flowrates condition Balanced flowrates condition is defined as an operation at which the flowrate of supercritical water is equal to the flowrate of the precursor streams (CCR-1, CCR-3, CCR-4, CCR-6, and CJM-1, CJM-3, CJM-4, CJM-6 in Table 1). On the other hand, if the supercritical water flowrate is not equal to that of the precursor streams, it is called an unbalanced flowrates condition (CCR-2, CJM-2, CCR-5, CJM-5, CCR-7 and CJM-7 in Table 1). The temperature variations along the z direction in both reactors between different inlet temperatures of supercritical water were simulated, It indicates that the simulated temperatures along the z direction at y = 1.7mm in a counter-current reactor caught the temperature bump feature near the exit region of the inner pipe, while the predicted temperatures at y = 1.9 mm reproduced experimental data in other regions. Taking into account of the possible variation of the thermocouples tips in x-y plane being 2mm, the simulated temperatures can be considered as in good agreement with measurements. For the confined jet mixer, the simulated temperatures with an inlet temperature of supercritical water of 350 oC showed a decrease after the exit of the inner pipe, which may be due to that the axial velocity of the supercritical water stream is low, hence only producing a weak recirculation zone near the exit region. At higher inlet temperatures of supercritical water, the simulated temperature decrease became much less noticeable as the higher momentum of the supercritical water jet produced a stronger recirculation zone. Overall, the predicted temperatures along the z direction at y = 1 and 1.2 mm are in
877

The axial velocity, temperature and tracer concentration along the central line of the reactors under three inlet temperatures of supercritical water with balanced flowrates (not shown here) indicated that the simulated temperatures along the z axis in the counter-current reactor produced much lower exit temperatures from the inner pipe. With an inlet temperature of 350oC, the exit temperature from the inner pipe for the counter-current reactor is 256oC, while the corresponding exit temperature for the confined jet mixer is 325oC, which is about 69oC higher. Similarly, the exit temperatures difference between the two reactors at supercritical water inlet temperatures of 400 and 450oC are about 10 and 27oC, respectively. This is due to that the counter-current reactor has a longer insertion length of the inner pipe. The insertion length for the counter-current reactor is 73.5 mm from central line of the outlet flow to the exit of the inner pipe, while the distance between the central line of the inlet pipe for precursor streams and the exit of the inner pipe is only 13 mm. The longer length in countercurrent reactor produces larger heat transfer area, hence more heat was transferred from the supercritical water to the fluid in the annual section, and resulting in higher temperature reduction in inner pipe. Accordingly, the axial velocities at the exit of the inner pipe for counter-current reactor were reduced to about 76%, 41% and 60% of their inlet axial velocities at supercritical water inlet temperatures of 350, 400 and 450oC, respectively. However, the corresponding axial exit velocities for confined jet mixer were only reduced to about 91%, 79% and 84% of their inlet axial velocities. The higher temperature and axial velocity at the exit of the inner pipe for confined jet led to fast decay of tracer concentration, hence better mixing with this reactor configuration. 5.4 Effect of inlet flowrate of s.c. water at balanced flowrates The simulated temperatures along the z direction for both reactors matched the experimental data. For the countercurrent reactor at lower inlet flowrate of supercritical water (CCR-3), temperature difference in the annual section (between z = 0 and 73.5mm) is about 60oC which is about 2

8th IFAC Symposium on Advanced Control of Chemical Processes Furama Riverfront, Singapore, July 10-13, 2012

times of the value for CCR-4. For the confined jet mixer at lower inlet flowrate of supercritical water stream (CJM-3), the temperature jumped from about 187oC at the exit to about 307oC within 3 mm after the exit, while for the higher inlet flowrate (CJM-4), this jump of temperature happened in the annual section before the exit and within about 1 mm. This indicates that CJM-3 has a weaker and smaller recirculation zone near the exit region of the supercritical water pipe. Temperature, axial velocity and tracer concentration along the central line for both reactors with a supercritical water inlet temperature of 400oC under two different balanced flowrates were also obtained. It was found that the lower inlet flowrate of supercritical water for the counter-current reactor (CCR-3) produced about 50oC lower exit temperature at the supercritical water exit. Similarly, the corresponding axial velocity of supercritical water for CCR-3 is reduced to only 21% of the inlet axial velocity, comparing to a value of 45% for CCR-4. Therefore, the recirculation zone generated with CCR-3 is smaller and weaker than that of CCR-4. Accordingly, the tracer concentration along the central line reduced faster for CCR-4 after the exit, indicating that CCR-4 has better mixing performance. For the confined jet mixer, central line temperatures and axial velocities for both CJM-3 and CJM-4 at the exit point are higher than those from the counter-current reactor, hence better mixing for confined jet. 5.5 Inlet temperatures of s.c water with unbalanced flowrates The temperature variations along the z direction in the counter-current reactor and the confined jet mixer between three different inlet temperatures of supercritical water stream were simulated under unbalanced flowrate conditions. For three different inlet temperatures of supercritical water, the predicted temperature along the z direction well duplicated the measured data, which is due to that the flowrate of supercritical water is 2.5 times of the precursor stream, hence the effect of heat loss through the annual wall becoming less important and the recirculation zone maintaining strong. The axial velocity, temperature and tracer concentration along the central line under three different inlet temperatures of supercritical water with unbalanced flowrates were obtained. As the jet of supercritical water with 2.5 times flowrate comparing to the precursor stream, the supercritical water jet dominated the performance near the exit region of the supercritical water pipe, hence the recirculation zone surrounding the jet. Results show that for both reactors with three different inlet temperatures, the predicted temperatures along central line are decreased at similar scale from the inlet to the exit point of the supercritical water. The decreases for the counter-current reactors (CCR-2, CCR-5, CCR-7) were slightly faster, which gave slightly lower axial velocity values at the exit point. Therefore, the confined jet mixer produced a better mixing performance. 5.6 Comparisons of balanced and unbalanced flowrates The temperature variations along the z direction in the counter-current reactor and the confined jet mixer between balanced and unbalanced flowrates, with an inlet temperature of supercritical water stream being 400oC, were simulated. It was found that the runs with unbalanced inlet flowrate for both reactors generated good agreement between prediction and measurement. The runs with unbalanced flowrate had
878

higher exit temperature at the exit point of the supercritical water pipe due to the higher inlet flowrate, and also higher temperature at the outlet of the final product stream. The higher exit temperatures also enhanced the recirculation zone, hence reduced the temperature variation near the exit region. Temperature, axial velocity and tracer concentration along the central line for both reactors with a supercritical water inlet temperature of 400oC under balanced and unbalanced flowrates were also simulated. With 2.5 times of balanced supercritical water flowrate and half of the balanced precursor stream flowrate, the unbalanced flowrates for both reactors produced higher exit temperatures of 3~5 oC at the exit point of the supercritical water stream. The axial exit velocity at the exit point with unbalanced flowrate for the counter-current reactor (CCR-5) is about 25% higher than that with the balanced flowrate (CCR-4). Similarly, for the confined jet mixer, the unbalanced flowrate case (CJM-5) has an axial exit velocity of 8% higher than the balanced flowrate case (CJM-4). By comparing the simulated results of two reactors, the exit temperatures and axial velocities for both balanced and unbalanced flowrates are higher for the confined jet mixer than the counter-current reactor. The predicted results with both balanced and unbalanced flowrates showed that the tracer concentration along central line for the confined jet mixer has decreased faster than the counter-current reactor. However, the tracer concentration obtained along central line with unbalanced flowrate showed a slower decay comparing to balanced flowrate. 6. FINAL REMARKS CFD models were developed for the study of the fluid flow and heat transfer patterns, as well as the mixing behaviour in counter-current and confined jet reactors used in continuous hydrothermal flow synthesis of nanomaterials. The velocity, temperature, and mixing in both reactor configurations were simulated and compared under various operating conditions. The predicted temperatures along the z direction in both reactors under different operating are in good agreement with experimental data. The study provides useful information for improvement of reactor design. For the counter-current reactor, the much lower exit temperatures, hence exit axial velocities, at the exit point of the inner pipe due to the heat loss through the pipe wall into the annual section were identified as one of the major factors affecting reactor performance. Optimising the insertion length of the inner pipe in the existing reactor configuration is likely to improve performance. Furthermore, the larger diameters used also caused less effective mixing, hence requiring further studies on the effect of the pipe diameters on performance. For the confined jet mixer, the positions of the inner pipe, pipe diameter, and operating conditions, as well as the use of the information to for control will be investigated in future. ACKNOWLEDGEMENT Support from EPSRC (EP/E040624/1 for Leeds and EP/E040551/1 for UCL) is acknowledged. Thank are also due to Johnson Matthey Plc, Corin Group Plc, Resource Efficiency KTN, Coates Lorilleux Ltd, AMR Technologies, Malvern Instruments Ltd and Nanoforce Technology Ltd.

8th IFAC Symposium on Advanced Control of Chemical Processes Furama Riverfront, Singapore, July 10-13, 2012

References 2009. National Institute of Standard and Technology [Online]. www.nist.gov. [Accessed Nov 2011]. 2010. ANSYS Fluent Package www.fluent.co.uk. ADSCHIRI, T., KANAZAWA, K. & ARAI, K. 1992. Rapid and continuous hydrothermal synthesis of boehmite particles in sub and supercritical water. The Journal of the American Ceramic Society, 75, 2615-2618. AIMABLE, A., MUHR, H., GENTRIC, C., BERNARD, F., LE CRAS, F. & AYMES, D. 2009. Continuous hydrothermal synthesis of inorganic nanopowders in supercritical water: Towards a better control of the process. Powder Technology, 190, 99-106. BOLDRIN, P., HEBB, A. K., CHAUDHRY, A. A., OTLEY, L., THIEBAUT, B., BISHOP, P. & DARR, J. A. 2006. Direct Synthesis of Nanosized NiCO2O4 Spinel and Related Compounds via Continuous Hydrothermal Synthesis Methods. Industrial and Engineering Chemistry Research, 46, 4830-4838. DEMOISSON, F., ARIANE, M., LEYBROS, A., MUHR, H. & BERNARD, F. 2011. Design of a reactor operating in supercritical water conditions using CFD simulations. Examples of synthesized nanomaterials. Journal of Supercritical Fluids, in press. KAWASAKI, S. I., SUE, K., OOKAWARA, R., WAKASHIMA, Y., SUZUKI, A., HAKUTA, Y. & ARAI, K. 2010. Engineering study of continuous supercritical hydrothermal method using a T-shaped mixer: Experimental synthesis of NiO nanoparticles and CFD simulation. Journal of Supercritical Fluids, 54, 96102. LAUNDER, B. E., REECE, G. J. & RODI, W. 1975. Peogress in the development of a Reynolds-stress turbulence closure. Journal of Fluid Mechanics, 68, 537566. LESTER, E., BLOOD, P., DENYER, J., GIDDINGS, D., AZZOPARDI, B. & POLIAKOFF, M. 2006. Reaction engineering: The supercritical water hydrothermal synthesis of nano-particles. Journal of Supercritical Fluids, 37, 209-214. LIU, H. Q. & MACEDO, E. A. 1995. Accurate correlations for the selfdiffusion coefficients of CO2, CH4, C2H4, H2O, and D2O over wide ranges of temperature and pressure. Journal of Supercritical Fluids, 8, 310-317. LIU, Y. & FOX, R. O. 2006. CFD predictions for chemical processing in a confined impinging-jets reactor. AIChE Journal, 52, 731-744. MA, C. Y., TIGHE, C. J., GRUAR, R. I., MAHMUD, T., DARR, J. A. & WANG, X. Z. 2011. Numerical modelling of hydrothermal fluid flow and heat transfer in a tubular heat exchanger under near critical conditions. Journal of Supercritical Fluids, 57, 236-246. MOUSSIERE, S., JOUSSOT-DUBIEN, C., GUICHARDON, P., BOUTIN, O., TURC, H.-A., ROUBAUD, A. & FOURNEL, B. 2007. Modelling of heat transfer and hydrodynamic with two kinetics approaches during supercritical water oxidation process. Journal of Supercritical Fluids, 43, 324-332. NARAYANANA, C., FROUZAKISA, C., BOULOUCHOSA, K., PRKOPSKY, K., WELLIG, B. & RUDOLF VON ROHRB, P. 2008. Numerical modelling
879

of a supercritical water oxidation reactor containing a hydrothermal flame. Journal of Supercritical Fluids, 46, 149-155. SIERRA-PALLARES, J., DANIELEL.MARCHISIO, D. L., ALONSO, E., PARRA-SANTOS, M. T., CASTRO, F. & COCERO, M. J. 2011. Quantification of mixing efficiency in turbulent supercritical water hydrothermal reactors Chemical Engineering Science, 66, 1576-1589. TIGHE, C. J., GRUAR, R. I., MA, C. Y., MAHMUD, T., WANG, X. Z. & DARR, J. A. 2012. Mixing and heat transfer in continuous hydrothermal flow synthesis reactors; An in-situ investigation of counter-current mixing. Journal of Supercritical Fluids, 62, 165-172. WAGNER, W. & PRU, A. 2002. The IAPWS Formulation 1995 for the Thermodynamic Properties of Ordinary Water Substance for General and Scientific Use. Journal of Physical and Chemical Reference Data, 31, 387-535. WOO, X. Y., TAN, R. B. H. & BRAATZ, R. D. 2009. Modeling and computational fluid dynamics-population balance equation-micromixing simulation of impinging jet crystallizers. Crystal Growth and Design, 9, 154-164.

You might also like