Вы находитесь на странице: 1из 8

Sustainable Maritime Transportation and Exploitation of Sea Resources Rizzuto & Guedes Soares (eds) 2012 Taylor &

p; Francis Group, London, ISBN 978-0-415-62081-9

A numerical investigation of exhaust smoke-superstructure interaction on a naval ship


S. Ergin, Y. Paral & E. Dobrucal
Istanbul Technical University, Department of Naval Architecture and Marine Engineering, Maslak, Istanbul, Turkey

ABSTRACT: The exhaust smoke-superstructure interaction for a generic frigate is investigated numerically. The frigate was driven by a CODOG system. The k- model is adopted for turbulent closure, and the governing equations in three dimensions are solved using a finite volume technique. The computations were performed for different yaw angles, efflux velocities and temperatures of the exhaust smoke. The cases with diesel engines and gas turbines are considered. The calculated streamlines, temperature contours and smoke concentrations are presented and discussed. Furthermore, the detailed predictions are compared with the available experimental measurements. A good agreement between the predictions and experiments is obtained. The study has demonstrated that computational fluid dynamics is a powerful tool to study the problem of exhaust smokesuperstructure interaction on ships. 1 INTRODUCTION of exhaust smoke dispersion from the ship stacks is extremely complicated task. Recently, a comprehensive review on the numerical modeling of exhaust smoke dispersion from ships is given by Kulkarni et al. (2011) and Mishra et al. (2010). Traditionally, the exhaust smoke dispersion has been investigated using scale models in wind tunnel (see, for example, Nolan 1946, Acker 1952, Isyunov et al., 1979 & Kulkarni et al., 2005). However, these experimental studies are expensive, lengthy and time consuming. In this paper, the exhaust smoke dispersion from a generic frigate and its interaction with rest of the superstructure are investigated numerically. The perspective view of the frigate model is shown in Figure 1. The frigate is driven by a combined diesel or gas turbine, CODOG system. The k- model is adopted for turbulent closure, and the

Nowadays, the understanding of the exhaust behavior of ship stacks has become quite important in the naval ship design due to advances in superstructure electronics (radar and antenna) and heat seeking missiles. The predictions of velocity and temperature fields of the ship exhaust plume from the stack are of vital importance for positioning and arranging of various superstructure electronics, weapons, gas turbine intake and ventilation intakes in the superstructure with the minimum interference with the hot exhaust smoke from the stack. However, all modern naval ships tend to favor short stacks and tall mast to house electronics. This makes them prone to the problem of smoke nuisance where hot exhaust gas gets entrapped into turbulent wake of the superstructure and deteriorates the performance of electronics, weapons, sensors, increase the ships infrared signature and hampers the internal ventilation and gas turbine intakes. This is also problem for the operation of helicopters from ships (see, for example, Park et al., 2011, Mishra et al., 2010, Ergin et al., 2010, Syms 2008, Kulkarni et al., 2005, 2007, Fitzgerald, 1986). The dispersion of exhaust smoke is affected by a large number of parameters such as efflux velocity and temperature of smoke, level of turbulence, wind velocity and direction, geometry of the structures on ships deck etc. (see for example, Baham et al., 1977; Fitzgerald, 1986, Jin et al., 2001, Park et al., 2011, Heywood, 1988). Therefore, the prediction

Figure 1.

Perspective view of the frigate model.

109

governing equations in three dimensions are solved using a finite volume technique. The computations were performed for different yaw angles, efflux velocities and temperatures of the exhaust smoke. However, the location, height and shape of the stack are kept the same. The calculated streamlines, temperature contours and smoke concentrations are presented and discussed. Furthermore, the predictions are compared with the available experimental measurements. A good agreement between the predictions and experiments is obtained. The study has demonstrated that computational fluid dynamics is a powerful tool to study the problem of exhaust smokesuperstructure interaction on ships and is capable of providing a means of visualising the dispersion of the exhaust smoke under different operating conditions very early in the design spiral of a ship. 2 PHYSICAL MODEL The exhaust smoke-superstructure interaction for a generic frigate is investigated numerically. Figure 1 shows the perspective view of the frigate model. The frigate is driven by a combined diesel or gas turbine, CODOG system. The over-all length, breadth and height of the frigate are 138 m, 18.5 m and 10 m, respectively. As seen from Figure 1, there is an APAR (Active Phased Array Radar) tower on the fore and 3-D radar on the aft part of the superstructure. A helipad is located at the back end of the superstructure. The exhaust stack is placed at about the centre of the superstructure. 3 MATHEMATICAL MODEL The numerical analysis is based on the timeaveraged equations describing the conservation of mass, momentum and energy, and because of the turbulence model, the equations governing the transport of turbulence kinetic energy k and its dissipation rate . It is assumed that the fluid is incompressible with constant thermophysical properties. The governing equations for three-dimensional and turbulent flow, in a Cartesian coordinate system and using tensor notation, can be written as Continuity equation: ( u j ) = 0 x j Momentum equation: ui p ( u j ui ) = eff f x j x j x j xi x (2) (1)

Energy equation: T ( u jT ) = T ,e + gi eff x j x j x j Turbulent kinetic energy: k ( u j k ) = k ,e + P eff x j x j x j Dissipation rate: 2 ( u j ) = (5) ,eff + C1P C2 f x j x j x j k k The effective viscosity in the above equations are defined as (4) (3)

eff = + t f

(6)

Figure 2.

Structure of the computational grid.

Figure 3. Comparison of the calculations with experiments for K = 2. a) Calculated streamlines, b) Experimental results (Kulkarni et al., 2005).

110

The turbulent viscosity is obtained from

k2

(7)

where C is a constant. In equations (3), (4) and (5), is the diffusion coefficient given by eff /T, eff /k and eff /, respectively. The values of the constants appearing in the above equations are C1 = 1.44, C2 = 1.92, C = 0.09, T = 1.0, k = 1.0, and = 1.3. These values have been used for many forced convection studies. The term P in equations (4) and (5) represents the generation of turbulence energy given by P = t ui x j ui u j + x j xi (8)

NUMERICAL MODEL AND BOUNDARY CONDITIONS

The finite volume method is employed to obtain the numerical solution of the governing equations

with using unstructured grid. The computational domain is a box shape volume that includes ships body above waterline. The computational domain is divided into a set of tetrahedral cells. Figure 2 shows the typical structure of the grid used in the computations. The Drichlet boundary condition is applied on the inlet of the box. While on the exit boundary, the Neumann boundary condition, i.e., the derivative of the solution was applied. The computational domain was large enough to minimize the inlet boundary influence on the mixing where the model ship placed. The calculations were carried out with 714700 cells. The grid points were distributed non-uniformly with higher concentration of grids closer to the walls, see Figure 2. The effect of grid size on the results was investigated for each case. Convective terms of momentum equations are discretized using the hybrid differencing scheme. Diffusive terms are discretized by the central differencing scheme. The derivation of pressure is based on the SIMPLE algorithm. The details of the discretisation and solution procedures are given in Malalasekara (1995), Patankar (1980), Ergin et al. (2005, 2010) and Ref. (ANYSYS 12.1, CFX),

a) K=1

c) K=3

b) K=2

d) K=4

Figure 4.

Streamlines and temperature distribution for the case with Texh = 300 K, = 10.

111

which also describes the computer code used in the present work. In the computations, four different velocity ratios (ratio of exhaust velocity to relative wind velocity), K = 1, 2, 3 and 4 are considered. At the exit of the stack, the constant exhaust gas temperatures of Texh = 15C, 200C, 300C and 400C are used as the boundary conditions. Six different yaw angles (direction of upcoming air flow), = 0, 5, 10, 15, 20 and 30 are considered. Furthermore, the cases of diesel engines on and gas turbines on are considered. The wind tunnel experiments of Kulkarni et al. (2005) were used to validate the numerical model used in the study. As can be seen from Figure 3, the agreement between the simulations and experiments are good.
a) =15.0

RESULTS AND DISCUSSIONS

The results of the numerical study are presented in Figures 48 for different velocity ratios (K), yaw angles () and exhaust gas temperatures at the exit of the stack (Texh). Figure 4a-d show the effects of velocity ratio, K on the dispersion of exhaust gases for Texh = 300C and = 10, when the main diesel engines are on. As can be seen from the streamlines in Figure 4a-c, the plume is directed towards the helicopter deck at the back end of the superstructure. Since the exhaust gas cannot rise with its momentum and buoyancy forces, and it cannot escape from the turbulent zone around superstructure. Therefore, the downwash phenomena occur. The results in Figure 4a-d show that the exhaust plume rises more as the velocity ratio increases. Figure 5 shows the effect of yaw angle, on the dispersion of the exhaust gases. The results show that the yaw angle up to = 10, the exhaust gases can rise (see, Figure 4) and no downwash phenomena occurs. However, when the yaw angle = 10, the exhaust gases directed to the back part of the superstructure, and, consequently, the downwash phenomena occurs. As can be seen from Figure 4 and 5, the exhaust plume rises to higher levels while the yaw angle increases from = 10 to 30. Figures 68 show the isosurfaces of the exhaust gas concentrations for different velocity ratios (K), yaw angles () and exhaust gas temperatures at the exit of the stack, Texh. The isosurface presents% 1 concentration value of the exhaust gas given at the stack exit. The concentration values are presented in Figure 6 for the velocity ratio values, K = 14 with Texh = 300 and = 10. As the velocity ratio, K increases, the isosurface area increases. It means that the higher concentration values are on the

b) =20.0

c) =30.0

Figure 5. Streamlines and temperature distributions for the case with K = 1 and Texh = 300.

112

back end of the superstructure including helicopter deck. When the yaw angle increases, the area occupied by the high concentration values first increase, and then decrease (see, Figures 6 and 7). The effect of exhaust gas temperatures at the exit of the stack on the plume rise and concentration

distribution can be seen in Figure 8. The results in the Figure are presented for the yaw angle, = 0 and velocity ratio K = 1. It can be seen that the plume and high concentration region rise more with increasing temperature. On the other hand, the size of the high concentration region decreases with increasing temperature.

a) K=1

a) =5

b) K=2

a) =15

c) K=3

a) =20

d) K=4
Figure 6. Exhaust gas concentration distribution (the isosurface presents % 1 concentration of the plume at the stack outlet) with Texh = 300 and = 10.

a) =30
Figure 7. Exhaust gas concentration distribution (the isosurface presents % 1 concentration of the plume at the stack outlet) with Texh = 300 and K = 1.

113

a) T = 15C

b) T = 200C

momentum equations have been solved numerically with k- turbulence model using finite volume method. It is showed that the dispersion of smoke is affected by efflux velocity, temperature, and turbulence, wind velocity and direction and geometry of the superstructure. The results show that the yaw angle larger than = 10 and the velocity ratio K = 2 should be maintained to avoid downwash phenomena. It is also found that the effect of the buoyancy forces on the plume rise is less significant in comparison with the momentum. The study has demonstrated that computational fluid dynamics is a powerful tool to study the problem of exhaust smokesuperstructure interaction on ships during the early stages of ship design. The computational fluid dynamics can also be used to solve the problems related with exhaust smokesuperstructure interaction on the already available ships. The study provides a further understanding of the exhaust smoke-superstructure interaction for naval ships. REFERENCES
Acker, H.C. 1952. Stack Design to Avoid Smoke Nuisance, SNAME Transactions, 60, 566593. ANSYS 12.1, CFX. Baham, G.J. & McCallum, D. 1977. Stack Design Technology For Naval and Merchant Ships, SNAME Transactions, Vol. 85, 324349. Ergin, S. & Ota, M. 2005. A Study of the Effect of Duct Width on Fully Developed Turbulent Flow Characteristics in a Corrugated Duct, Heat Transfer Engineering, Vol. 26, No.2, s. 5462. Ergin, S. & Paral, Y. 2010. Numerical Study of Interaction between the Exhaust Gases and Superstructure of a Naval Ship, Gemi ve Deniz Teknolojisi Dergisi, 185 (in Turkish). Fitzegerald, M.P. 1986. A Method to Predict Stack Performance, Naval Engineers Journal, 541. Heywood, J.B. 1988. Internal Combustion Engine Fundamentals, McGraw-Hill. Isyumov, N. & Tanaka, H. 1979. Wind-tunnel Modelling of Stack Gas Dispersion Difficulties and Approximations, Proc. of the fifth Int. Conf. on Wind Engineering, Colorado, USA. Jin, E., Yoon, J. & Kim, Y. 2001. A CFD Based Parametric Study on the Smoke Behavior of a Typical Merchant Ship, Practical Design of Ships and Other Floating Structures, 459465. Kulkarni, P.R., Singh, S.N. & Seshadri, V. 2005. Flow Visualisation Studies of Exhaust SmokeSuperstructure Interaction on Naval Ships, Naval Eng J, ASNE, 117, 4156. Kulkarni, P.R., Singh, S.N. & Seshadri, V. 2007. Parametric Studies of Exhaust Smoke-Superstructure

c) T = 300C

d) T = 400C
Figure 8. Exhaust gas concentration distribution (the isosurface presents % 1 concentration of the plume at the stack outlet) with = 0 and K = 1.

6 CONCLUSIONS The exhaust smoke-superstructure interaction for a generic frigate is investigated through several calculations. The conservation of mass and

114

Interaction on a Naval Ship Using CFD, Computers & Fluids 36, 794816. Malalasekera, W. & Versteeg, H.K. 1995. An Introduction to Computational Fluid Dynamics the Finite Volume Method, Longman Scientific and Technical. Mishra, D.P. & Dash, S.K. 2010. Numerical Investigation of Air Suction Through the Louvers of a Funnel due to High Velocity Air Jets, Computers & Fluids, 39, 15971608. Nolan, R.W. 1946. Design of Stacks to Minimize Smoke Nuisance. Trans SNAME, 54, 4284.

Park, S., Heo, J., Yu, B.S., Rhee & S.H. 2011. Computational analysis of ships exhaust-gas flow and its application for antenna location, Applied Thermal Engineering, doi:10.1016/j.applthermaleng. 2011.02.011 (in press). Patankar, S.V. Numerical Heat Transfer and Fluid Flow, McGraw-Hill, New York, 1980. Syms, G.F. 2008. Simulation of simplified-frigate airwakes using lattice-Boltzman Method, Journal of Wind Engineering and Industrial Aerodynamics, 96, 11971206.

115

Вам также может понравиться