You are on page 1of 16

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/268276921

Experimental and Numerical investigation of the Downburst wind profiles on


different surfaces

Article

CITATIONS READS

0 126

3 authors, including:

Kalyan Kumar Das K. P. Sinhamahapatra


Assam Engineering College Indian Institute of Technology Kharagpur
7 PUBLICATIONS   13 CITATIONS    43 PUBLICATIONS   368 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

low gravity dynamics View project

Aerodynamics and Aeroacoustics of Supersonic Jets View project

All content following this page was uploaded by K. P. Sinhamahapatra on 09 December 2014.

The user has requested enhancement of the downloaded file.


K.K.Das et al. / International Journal of Engineering Science and Technology (IJEST)

Experimental and Numerical investigation


of the Downburst wind profiles on different
surfaces

K.K.Das1*, A.K.Ghosh2, K.P.Sinhamahapatra3

1*
Department of Mechanical Engineering, Assam Engineering College,Guwahati, INDIA
23
Department of Aerospace Engineering, Indian Institute of Technology Kharagpur, INDIA
*
Corresponding Author: e-mail: kkdas1971@gmail.com, Tel +91-09864405087

Abstract

Severe thunderstorms are important weather phenomenon which impact on various facets of national activity

like civil and defense operation, particularly aviation, space vehicle launching, agriculture in addition to its

damage potential to life and properties. Experimental and numerical simulation studies on thunderstorm

downburst have been reported by many researchers during the past two decades. Axisymmetric numerical code

has been developed to simulate the downburst wind by using the vorticity-stream formulation, with LES for the

turbulence. In addition a microburst simulator has been fabricated with a 165 mm diameter nozzle to generate

experimental data for the translational downburst.

Keywords: Microburst, Experimental simulation, Ring vortex, Macro-flow dynamics, Partial slip.

1.0 Introduction

The famous atmospheric scientist Fujita(1981) has observed and studied the flow due to downburst impacting

on the ground and spreading outward in the different directions. He classified downburst as either microburst or

macroburst depending on their horizontal extent of damage. For the complexity of the full scale phenomenon,

the physical simulation of the downburst is confined to the generic experiments of density currents impinging on

a wall. Alahyari and Longmire(1995), Lundgren et al.(1992), Cooper et al. (1993), Didden and Ho(1985),

Knowles and Myszko(1998) have studied experimental simulation of the downburst. Letchford and Chay(2002),

Chay and Letchford(2002) and Sengupta and Sarkar (2007) performed physical modelling to study the flow

ISSN : 0975-5462 Vol. 3 No. 6 June 2011 4668


K.K.Das et al. / International Journal of Engineering Science and Technology (IJEST)

field characteristics and pressure distribution the stationary and translational downburst. Numerical simulation

of the downburst is performed by Proctor(1988),Craft et al. (1993) ,Selvam and Homes(1992) Das et al.(2010).

Kim and Hangan(2006) and Sengupta and Sarkar (2007) simulated the downburst flow field with different

turbulence model using FLUENT software. The primary objective of this work is to investigate the velocity

profiles of the downburst wind on different surfaces.

2.0 Numerical Simulation

The two-dimensional incompressible Navier-Stokes equations in stream function-vorticity form are solved

numerically to simulate the axisymmetric impinging jet downburst problem. The LES technique is adopted to

model the turbulence. The Poisson equation for the stream function ψ is given by

V2ψ=-ξ ---- (1)

the vorticity transport equation in non-dimensional form is


 
  u 

 v

 
1
     sgs
 2 
 
 
t x y Re     
   
  
2  u  2     sgs     sgs     sgs
 v  2    2 (2)
      2  u  v    
Re  y x 2    x        
   y 2    x y  xy   

The vorticity transport equation (2) is normalized with the jet exit parameters. The unsteady vorticity transport

equation is parabolic, and is solved using the implicit ADI technique, whereas the Poisson equation for stream

function is elliptic and is solved by successive over-relaxation (SOR) method with a relaxation factor (ω) of

1.85. It is further assumed that the flow enters the computational domain with the jet exit velocity, where the

fluid is stationary at t = 0. The Smagorinsky constant (Cs) is taken as 0.15 for this CFD simulation.

2.1 Boundary Conditions

Figure 1 shows the two-dimensional computational domain and its boundaries schematically. The impermeable

no-slip boundary condition is imposed on the solid walls as shown in figure 4.1. For the outflow boundary BC,

is equated to zero and for the outflow boundary at CD, is assumed to be zero. The zero shear stress

condition is imposed on the slip wall DE. At the inflow boundary EF, the radial velocity is zero and the axial

velocity is equal to the velocity of the jet. Symmetry condition is imposed on the axis AF.

ISSN : 0975-5462 Vol. 3 No. 6 June 2011 4669


K.K.Das et al. / International Journal of Engineering Science and Technology (IJEST)

Figure 1: Computational Domain for the numerical simulation

3.0 Physical modeling

Physical simulation of the translational dry microbursts is done using the impinging jet model with a 165 mm

exit diameter pipe. Two 1.5 HP centrifugal blowers are used to generate the impinging jet. The dimension of the

wooden platform on which the jet impinges is 2.0 m × 2.0 m with roughness of 4.2 micron. The distance of the
jet from the impinging platform (H) can be varied between 125 mm and 400 mm using an adjustable frame to

change the value of H/Djet. The H/Djet ratios considered for this work are 1.0, 1.5, 2.0. Three jet velocities (Vjet)

of 10 m/s, 15 m/s and 20 m/s are used in the experiments. Experimental setup is shown in figure 2. Velocity is

measured using DANTEC 56C17 CTA probe with traverse system and CTA software.

Three different surfaces are considered in the present simulations as shown in table 1. Surface Roughness is

measured by Taylor-Hobson tally surf. Figure 3 shows the photographs of vortex formation on the three surfaces

at Reynolds number of 2.2105 and H/Djet =1.0. Photographs are taken at non-dimensional time 1.0.

ISSN : 0975-5462 Vol. 3 No. 6 June 2011 4670


K.K.Das et al. / International Journal of Engineering Science and Technology (IJEST)

Table 1

Sl. Type of test surface Roughness in


number micron
1 Wood 4.2
2 Ply 2.6
3 Perspex 0.72

(a) (b)
.

Figure 2 Physical simulator of the thunderstorm microburst fabricated at IIT Kharagpur

Figure 3 Photographs showing formation of vortices on three different surfaces.

ISSN : 0975-5462 Vol. 3 No. 6 June 2011 4671


K.K.Das et al. / International Journal of Engineering Science and Technology (IJEST)

Figure 4 Flow visualization of the simulated travelling downburst from the physical simulator

From figure 3 it is found that the primary vortex forms at a longer distance from the impinging plate when the

surface is rougher. Due to thicker boundary layer in rough surface the vortex is lifted away from the plate and as

a consequence less amount of smoke can be seen close to the plate for rougher surface. The primary, secondary

ISSN : 0975-5462 Vol. 3 No. 6 June 2011 4672


K.K.Das et al. / International Journal of Engineering Science and Technology (IJEST)

and tertiary vortices can be seen in the flow field on each of the surfaces. To investigate the effects of surface

slipping numerically the concept of partial slip is introduced. Slipping (S) is mathematically defined as,

where is the velocity on the test surface and is the velocity at a neighboring node. For no-slip surface

S=0% and for free slip surface S=100%. For the numerical investigation three types of surfaces are considered

having, respectively 0%, 5% and 10% slipping. Figure 7 shows the velocity fields on these surfaces at the

Reynolds number of 2.2×105 and H/Djet =1.0

3.1 Velocity and Pressure measurements

A DANTEC 56C17 hot wire anemometer system is used to measure the velocity in the flow field. In addition, a

vane type digital anemometer is also used to measure the velocity at some locations in the flow field. To

determine the velocity profiles in the radial and axial directions, hot wire anemometer probe is placed in the

DANTEC traversing system as shown in figure 5. Pressure is measured using a PDCR23 pressure transducer

system with a scanivalve. To estimate the pressure on the impinging platform 300 pressure taps are placed on

the platform. Pressure taps are connected to the scanivalve through 1 mm diameter PVC tubing. A multi tube

manometer is also used to verify the pressure readings of the PDCR23 pressure transducer system.

(a)

ISSN : 0975-5462 Vol. 3 No. 6 June 2011 4673


K.K.Das et al. / International Journal of Engineering Science and Technology (IJEST)

(b) (c)

Figure 5 DANTEC CTA Probe with the Traversing system and the controller

3.2 Flow visualization

Flow visualisation of the impinging jet is done using a smoke generator and high speed cameras. Smoke

generator is connected to the inlet of the blower. Flow patterns at different jet velocities and plate locations are

photographed.

4.0 Results and Discussions

Some assumptions are made in the present numerical and experimental simulations of the dry downburst. It is

assumed that the buoyant acceleration characteristics of a natural downburst can be modeled by impinging jet.

Cross jets are used to generate secondary injection, which reduces the boundary layer effect of the jet wall,

which gives a better representation of the density driven flow in the natural downburst.

The spatial scale of the simulation is estimated based on the observations related to stationary microbursts made

by Hjelmfelt (1988). Hjelmfelt observed that the microburst typically had a diameter of 1.8 km and that the

maximum outflow winds occurred at approximately 1.5 km from the center of the descending column of air.

Based on this observation the geometric scaling factor in this study is about 1/10000. Also, it is found from the

preliminary tests performed in the physical simulation that the model produces peak radial wind of nearly 32m/s

compared to a maximum velocity of nearly 60m/s in natural downburst (Fujita, 1981). Therefore the velocity

scaling factor is 1/2. Combining the geometric and velocity scaling factors lead to the time scale of 1/5000 for

this study.

ISSN : 0975-5462 Vol. 3 No. 6 June 2011 4674


K.K.Das et al. / International Journal of Engineering Science and Technology (IJEST)

Four distinct major vortices are observed within the downburst flow field: a primary, an intermediate, a trailing

and a counter-rotating secondary vortex as seen in figure 6. Several transient vortices are also observed near the

test surface immediately after the impact of the jet on the test surface as seen in figure 6. The secondary vortex

is generated at x = 1.8 – 3.5. From these figures it is observed that the axial location of the primary vortex is

practically unaltered due to variation in plate separation ratio, but the variation in radial location is significant.

The primary and secondary vortices form at a larger radial distance from the point of impact when the plate

separation ratio is smaller. Similar results are also observed for other plate separations and Reynolds numbers. 

Figure 6 Vortex formations in the travelling downburst from the numerical simulation

Figure 7 shows the velocity fields of the downburst wind on different surfaces having patilal slips 0%, 5% and

10% from the numerical simulation. It can be observed from figures 7(a), (b) & (c) that the location of the

primary vortex is different in the three cases. The primary vortex forms at longer radial and axial distance

respectively from the jet axis and test surface for rougher surface. Similar results are obtained for other

Reynolds numbers and plate separations as well. The radial velocity profiles at x=1.2 on the three surfaces at

Reynolds number of 2.2105 and H/Djet =2.0 are shown in figure 8. It shows that the maximum radial velocity

increases at higher value of slip. In addition the location of the maximum velocity point is closer to the ground

at greater value of slip. Considerable change in radial velocity is observed as the slip increases from 5% to 10%.

However, the radial velocity does not change significantly when the slip increases from 0 to 5%. The behavior is

ISSN : 0975-5462 Vol. 3 No. 6 June 2011 4675


K.K.Das et al. / International Journal of Engineering Science and Technology (IJEST)

attributed to the considerably thicker boundary layer at 10% slip and to the associated secondary and tertiary

vortices.

Figure 8 also reveals a very interesting trend in the velocity profile at S= 10% where the velocity is almost

constant above y=0.7. In the range of y=0.07 – 0 there is a sharp decay in radial velocity due to the formation of

primary vortex. Uniform radial velocity prevails beyond y=0.7. The velocity profiles for S=0% and S=10% are

significantly different due to reduced surface slip.

Figure 9 shows the radial velocity profiles at two locations at Reynolds number of 2.2105 and H/Djet =2.0 with

10% surfaces slip. Similar to no-slip surface, partial slip surfaces also show maximum radial velocity close to

the surface. However, the location of the maximum velocity point is different. The maximum radial velocity

point moves closer to the surface at higher slip value. Similar flow behavior is also seen at other Reynolds

numbers and plate separations. Figure 10(a) and (b) shows the ground pressure distribution obtained

respectively from the experimental and numerical simulations at Reynolds number 2.2105 and H/Djet =2.0.

Qualitatively identical results are obtained at the other Reynolds numbers and H/Djet values. Simulated results

clearly indicate that the ground pressure coefficient distribution is practically insensitive to surface slip.

The computed and experimental radial velocity profiles from the present simulations are compared with the field

observation data from the project NIMROD (Fujita, 1981) and empirical profiles due to Rajaratnam (1976) and

Wood et al. (2001). The comparison is presented in figure 11. The radial velocity in the figure is normalized

with the maximum radial velocity and the height is normalized with respect to the height at which radial

velocity falls to 50% of the maximum. The experimental and numerical results shown in figure 8 are for jet

velocity 30 m/s and H/Djet =1.0. The computed radial velocity profile matches closely with the experimental

data. The two-dimensional computation too agrees reasonably well. Very good agreement is also observed with

the full scale data from NIMROD (Fujita, 1981) as shown in figure 11.

ISSN : 0975-5462 Vol. 3 No. 6 June 2011 4676


K.K.Das et al. / International Journal of Engineering Science and Technology (IJEST)

(a)

(b)

ISSN : 0975-5462 Vol. 3 No. 6 June 2011 4677


K.K.Das et al. / International Journal of Engineering Science and Technology (IJEST)

(c)
Figure 7 Velocity fields of the downburs wind on different surfaces

 
Figure 8 Radial velocity profiles on different surfaces at x=1.2 from 2D numerical simulation.

ISSN : 0975-5462 Vol. 3 No. 6 June 2011 4678


K.K.Das et al. / International Journal of Engineering Science and Technology (IJEST)

Figure 9 Computed radial velocity profiles at different locations with 10 % surface slip

ISSN : 0975-5462 Vol. 3 No. 6 June 2011 4679


K.K.Das et al. / International Journal of Engineering Science and Technology (IJEST)

(a)

(b)

Figure 10 Ground pressure coefficient distribution at Reynolds number 2.2105 and H/Djet=2.0 for different surfaces: (a)
experimental; (b) numerical

5.0 Conclusion

A CFD code is developed to investigate the velocity and pressure profiles of the downburst wind on surfaces

having different roughness. Physical simulator is also fabricated for the validation of the code based on the

impinging jet model. Simulated downburst results are compared with the results from the NIMROD full scale

data.

ISSN : 0975-5462 Vol. 3 No. 6 June 2011 4680


K.K.Das et al. / International Journal of Engineering Science and Technology (IJEST)

The present study reveals the following facts regarding the travelling downburst flow field,

a. Five distinct vortices coexist in the simulated downburst flow field. The trailing and intermediate vortices form

before the jet impact and the primary, secondary and tertiary vortices form after the impact of the jet in the wall

jet region. Several transient vortices form near the test surface immediately after the jet impact which produces

tremendous wind shear near the ground.

b. Separation and reattachment of the transient vortices strengthen the primary vortex near the ground and

produces high velocity near the ground. The maximum velocity occurs at about 5 – 7% of Djet above the ground

and about 1.5Djet away from the jet axis.

c. Ground pressure coefficient distribution is independent of Reynolds number and jet separation ratio.

d. The magnitude of the maximum radial velocity increases for decreasing value of surface slip but the maximum

velocity occurs closer to the surface for higher value of slip.

Figure 11 Comparison with full scale data

6.0 Nomenclature

x Radial direction

y Axial direction

Vjet Jet velocity

u Radial velocity

ISSN : 0975-5462 Vol. 3 No. 6 June 2011 4681


K.K.Das et al. / International Journal of Engineering Science and Technology (IJEST)

v Axial velocity

Djet Diameter of the jet

H Distance of the jet from the impinging plate.

H/Djet Plate separation ratio, Cloud height for the full scale downburst.

7.0 References

[1] Alahyari, A., Longmire, E.K., 1995. Dynamics of experimentally simulated microbursts. AIAA J. 33 (11), 2128-2136.
[2] Chay, M.T., Letchford, C.W., 2002. Pressure distribution on a cube in a simulated thunderstorm downburst—Part A: stationary
downburst simulation. J. Wind Eng. Ind. Aerodyn. 90, 711-732.
[3] Cooper, D., Jackson, D.C., Launder, B.E., Liao, G.X., 1993. Impinging jet studies for turbulence model
[4] Assessment-I. Flow-field experiments. Int. J. Heat Mass Transfer 36 (10), 2675–2684.
[5] Craft, T.J., Graham, L.J.W., Launder, B.E., 1993. Impinging jet studies for turbulence model Assessment-II: an ex amination of
the performance of four turbulence models. Int. J. Heat Mass Transfer 36 (10), 2685–2697.
[6] Das K.K, Ghosh A.K., Sinhamahapatra K.P.,2010, Investigation of the axisymmetric microburst flow field, Journal of Wind and
Engg., Vol. 7 no. 1, Jan 2010, pp 1-15.
[7] Didden, N., Ho, C.M., 1985. Unsteady separation in a boundary layer produced by an impinging jet. J. Fluid
[8] Mech. 160, 235–256.
[9] Fujita, T.T., 1981. Tornadoes and downbursts in the context of generalized planetary scales. J. Atmos. Sci. 38, 1511–1534.
[10] Fletcher C.A.J 1987 Computational Techniques for the Fluid Dynamics(vol. 2) Springer-Verlag Publication
[11] Hjelmfelt, M.R., 1988. Structure and life cycle of micoburst outflows observed in Colorado. J. Appl. Met. 27, 1988, 900-927
[12] Holmes, J.D., Oliver, S.E., 2000. An empirical model of a downburst. Eng. Struct. 22, 1167–1172.
[13] Kim, J., Hangan, H., 2007. Numerical simulation of impinging jets with application to downbursts. J. Wind Eng. Ind. Aerodyn.
95, 279–298.
[14] Knowles, K., Myszko, M., 1998. Turbulence measurement in radial wall-jets. Exp. Thermal Fluid Sci. 17, 71–78.
[15] Letchford, C.W., Chay, M.T., 2002. Pressure distributions on a cube in a simulated thunderstorm downburst, Part B: moving
downburst observations. J. Wind Eng. Ind. Aerodyn. 90, 733–753.
[16] Lundgren, T.S., Yao, J., Mansour, N.N., 1992. Microburst modeling and scaling. J. Fluid Mech. 239, 461–488.
[17] Proctor, F.H., 1988. Numerical simulations of an isolated microburst. Part I: Dynamics and Structure. J. Atmos. Sci. 45, 3137–
3160
[18] Sakamota, S, Murakami S., Mochida A., 1993. Numerical study on flow past 2D square cylinder by Large Eddy Simulation
Comparison between 2D and 3D computations, J. Wind Eng. Ind. Aerodynamics 50 (1993) 61-68.
[19] Selvam, R.P., Holmes, J.D., 1992. Numerical simulation of thunderstorm downdrafts. J. Wind Eng. Ind. Aerodyn. 41–44, 2817–
2825
[20] Sengupta, A., Sarskar, P. P., 2007. Experimental measurement and numerical simulation of an impinging jet with application to
thunderstorm microburst winds. J. Wind Eng. Ind. Aerodyn. (2007)

ISSN : 0975-5462 Vol. 3 No. 6 June 2011 4682

View publication stats

You might also like